09.12.2012 Views

q 2006 by Taylor & Francis Group, LLC - Developers

q 2006 by Taylor & Francis Group, LLC - Developers

q 2006 by Taylor & Francis Group, LLC - Developers

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Dedication<br />

To the loving memory of my dad,<br />

Mustafa Amiji, and to my mom, Shirin Amiji


Foreword<br />

We are at the threshold of a new era in cancer treatment and diagnosis, brought about <strong>by</strong> the<br />

convergence of two disciplines—materials engineering and life sciences—that 30 years ago might<br />

have been difficult to envision. The product of this curious marriage, nanobiotechnology, is yielding<br />

many surprises and fostering many hopes in the drug-development space. Nanoparticles,<br />

engineered to exquisite precision using polymers, metals, lipids, and carbon, have been combined<br />

with molecular targeting, molecular imaging, and therapeutic techniques to create a powerful set of<br />

tools in the fight against cancer. The unique properties of nanomaterials enable selective drug<br />

delivery to tumors, novel treatment methods, intraoperative imaging guides to surgery, highly<br />

sensitive imaging agents for early tumor detection, and real-time monitoring of response to<br />

treatment.<br />

Using nanotechnology, it may be possible to leap over many of the hurdles of cancer drug<br />

delivery that have confounded conventional drugs. These hurdles include hindered access to the<br />

central nervous system through the blood–brain barrier, sequestration <strong>by</strong> the reticulo-endothelial<br />

system, inability to penetrate the interior of solid tumors, and overcoming multi-drug resistance<br />

mechanisms. These obstacles can be mitigated <strong>by</strong> manipulating the size, surface charge,<br />

hydrophilicity, and attached targeting ligands of a therapeutic nanoparticle.<br />

The novelty of using nanotechnology for medical applications presents its own challenges,<br />

similar in many ways to the challenges faced <strong>by</strong> the introduction of the first synthetic protein-based<br />

drugs. In the early 1980s, protein-based drugs offered an entirely new approach to targeting and<br />

treating disease. They could be “engineered” to be highly specific and even offered the capability of<br />

separating the targeting and therapeutic functions of a drug, such as when monoclonal antibodies<br />

are conjugated to cytotoxic agents. But monoclonal antibodies and recombinant protein-based<br />

drugs necessitated a different approach to lead optimization, metabolism, and toxicity screening<br />

from that of small-molecule drugs. Factors such as stability, immunogenicity, and species<br />

specificity had to be considered and tested in ways that were not familiar to the developers of<br />

traditional drugs.<br />

Nanotechnology-based drugs will catalyze a reevaluation of optimization, metabolism, and<br />

toxicity-screening protocols. Interactions between novel materials and biological pathways are<br />

largely unknown but are widely suspected to depend heavily on physical and chemical<br />

characteristics such as particle size, particle size distribution, surface area, surface chemistry<br />

(including charge and hydrophobicity), shape, and aggregation state—features not usually<br />

scrutinized for traditional drugs. Standard criteria for physicochemical characterization will have to<br />

be established before safety testing can yield interpretable and reproducible results.<br />

Would nanotechnology-based drugs be successful? All drugs have to run an obstacle course<br />

through physiological and biochemical barriers. For monoclonal antibody drugs, only a minute<br />

portion of an intravenously administered drug reaches its target. However, drugs based on<br />

nanomaterials, including nanoparticles, have unique properties that enable them to specifically bind<br />

to and penetrate solid tumors. Size and surface chemistries can be manipulated to facilitate<br />

extravasation through tumor vasculature, or the therapeutic agent can be encapsulated in polymer<br />

micelles or liposomes to prevent degradation and increase circulation half-life to improve the odds<br />

of it reaching the target.<br />

This level of functional engineering has not been available to either small-molecule or proteinbased<br />

drugs. For these drugs, modification of one characteristic, such as solubility or charge, can<br />

have dramatic effects on other essential characteristics, such as potency or target specificity.<br />

Nanoparticles, however, introduce a much higher degree of modularity and offer at least four<br />

advantages over the antibody conjugates: (1) the delivery of a larger therapeutic payload per target<br />

recognition event, enhancing potency; (2) the ability to carry multiple targeting agents, enhancing<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


selectivity; (3) the ability to carry multiple therapeutic agents, enabling targeted combination<br />

therapies; and (4) the ability to <strong>by</strong>pass physiological and biological barriers.<br />

We would all like to hasten the day when chemotherapy—the administration of non-specific,<br />

toxic anti-cancer agents—is relegated to medical history. Improvements in diagnostic screening<br />

and the development of drugs that target specific biological pathways have begun to turn the tide<br />

and contribute to a slow, yet consistent decline in death rates. While confirming that early diagnosis<br />

and targeted therapies are pointing us in the right direction, the progress is still incremental.<br />

Nanotechnology may be our best hope for overcoming many of the barriers faced <strong>by</strong> today’s drugs<br />

in the battle against cancer. The combined creative forces of engineering, chemistry, physics, and<br />

biology will beget new and hopefully transformative options to old, intractable problems.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Piotr Grodzinski, Ph.D.<br />

National Cancer Institute


Preface<br />

With parallel breakthroughs occurring in molecular biology and nanoscience/technology, the newly<br />

recognized research thrust on “nanomedicine” is expected to have a revolutionary impact on the<br />

future of healthcare. To advance nanotechnology research for cancer prevention, diagnosis, and<br />

treatment, the United States National Cancer Institute (NCI) established the Alliance for<br />

Nanotechnology in Cancer in September 2004 and has pledged $144.3 million in the next five years<br />

(for details, visit http://nano.cancer.gov). Among the approaches for exploiting developments in<br />

nanotechnology for cancer molecular medicine, nanoparticles offer some unique advantages as<br />

sensing, delivery, and image enhancement agents. Several varieties of nanoparticles are available,<br />

including polymeric conjugates and nanoparticles, micelles, dendrimers, liposomes, and<br />

nanoassemblies.<br />

This book focuses specifically on nanoscientific and nanotechnological strategies that are<br />

effective and promising for imaging and treatment of cancer. Among the various approaches<br />

considered, nanotechnology offers the best promise for targeted delivery of drugs and genes to the<br />

tumor site and alleviation of the side effects of chemotherapeutic agents. Multifunctional<br />

nanosystems offer tremendous opportunity for combining more than one drug or using drug and<br />

imaging agents. The expertise of world-renowned academic and industrial researchers is brought<br />

together here to provide a comprehensive treatise on this subject.<br />

The book is composed of thirty-eight chapters divided into seven sections that address the<br />

specific nanoplatforms used for imaging and delivery of therapeutic molecules. Section 1 focuses<br />

on the rationale and fundamental understanding of targeting strategies, including pharmacokinetic<br />

considerations for delivery to tumors in vivo, multifunctional nanotherapeutics, boron neutron<br />

capture therapy, and the discussion on nanotechnology characterization for cancer therapy, as well<br />

as guidance from the U.S. Food and Drug Administration on approval of nanotechnology products.<br />

Section 2 focuses on polymeric conjugates used for tumor-targeted imaging and delivery, including<br />

special consideration on the use of imaging to evaluate therapeutic efficacy. In Section 3, polymeric<br />

nanoparticle systems are discussed with emphasis on biodegradable, long-circulating nanoparticles<br />

for passive and active targeting. Section 4 focuses on polymeric micellar assemblies, where<br />

sophisticated chemistry is applied for the development of novel nanosystems that can provide<br />

efficient delivery to tumors. Many of the micellar delivery systems are undergoing clinical trials in<br />

Japan and other countries across the globe. Dendritic nanostructures used for cancer imaging and<br />

therapy are discussed in Section 5. Section 6 focuses on the oldest nanotechnology for cancer<br />

therapy—liposome-based delivery systems—with emphasis on surface modification to enhance<br />

target efficiency and temperature-responsive liposomes. Lastly, Section 7 focuses on other lipid<br />

nanosystems used for targeted delivery of cancer therapy, including nanoemulsions that can cross<br />

biological barriers, solid-lipid nanoparticles, lipoprotein nanoparticles, and DQAsomes for<br />

mitochondria-specific delivery.<br />

Words cannot adequately express my admiration and gratitude to all of the contributing authors.<br />

Each chapter is written <strong>by</strong> a world-renowned authority on the subject, and I am deeply grateful for<br />

their willingness to participate in this project. I am also extremely grateful to Dr. Piotr Grodzinski<br />

for providing the Foreword. Drs. Fredika Robertson and Mauro Ferrari have done a superb job in<br />

laying the foundation <strong>by</strong> providing a chapter entitled “Introduction and Rationale for<br />

Nanotechnology in Cancer.” I am grateful to Professor Kinam Park at Purdue University,<br />

Professor Robert Langer at MIT, and Professor Vladimir Torchilin at Northeastern University, who<br />

have been my mentors and collaborators, as well as many other researchers from academia and<br />

industry. Special thanks are due to the postdoctoral associates and graduate students in my<br />

laboratory at Northeastern University who have been the “soldiers in the trenches” in our quest to<br />

use nanotechnology for the targeted delivery of drugs and genes to solid tumors. Lastly, I am deeply<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


grateful to the wonderful people at <strong>Taylor</strong> & <strong>Francis</strong>-CRC Press, including Stephen Zollo, Patricia<br />

Roberson, and many others, who have made the concept of this book into reality.<br />

Any comments and constructive criticisms of the book can be sent to the editor at<br />

m.amiji@neu.edu.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Editor<br />

Dr. Mansoor M. Amiji received his undergraduate degree in pharmacy from Northeastern<br />

University in 1988 and his PhD in pharmaceutics from Purdue University in 1992. His areas of<br />

specialization include polymeric biomaterials, advanced drug delivery systems, and nanomedical<br />

technologies.<br />

Dr. Amiji’s research interests include the synthesis of novel polymeric materials for medical and<br />

pharmaceutical applications; surface modification of cationic polymers <strong>by</strong> the complexationinterpenetration<br />

method to develop biocompatible materials; the preparation and characterization<br />

of polymeric membranes and microcapsules with controlled permeability properties for medical<br />

and pharmaceutical applications; target-specific drug and vaccine delivery systems for<br />

gastrointestinal tract infections; localized delivery of cytotoxic and anti-angiogenic drugs for<br />

solid tumors in novel biodegradable polymeric nanoparticles; intracellular delivery systems for<br />

drugs and genes using target-specific, long-circulating, biodegradable polymeric nanoparticles; and<br />

gold and iron-gold core-shell nanoparticles for biosensing, imaging, and delivery applications. His<br />

research has received sustained funding from the National Institutes of Health (NIH), the National<br />

Science Foundation (NSF), foundations, and local industries.<br />

Dr. Amiji is Professor and Associate Chair of the Pharmaceutical Sciences Department and Co-<br />

Director of the Northeastern University Nanomedicine Education and Research Consortium<br />

(NERC). The NERC oversees a doctoral training grant in nanomedicine science and technology<br />

that was co-funded <strong>by</strong> the NIH and NSF. He has two published books, Applied Physical Pharmacy<br />

and Polymeric Gene Delivery: Principles and Applications, along with numerous manuscript<br />

publications. He has also received a number of awards, including the 2003 Eurand Award for<br />

Innovative Oral Drug Delivery Research, Third Prize.<br />

Dr. Amiji has supervised the research efforts of over 50 postdoctoral associates, doctoral and<br />

master’s level graduate students, and undergraduate honors students over the last 13 years. His<br />

teaching responsibilities include the Doctor of Pharmacy (PharmD) program and graduate<br />

programs (MS and PhD) in pharmaceutical sciences, biotechnology, and nanomedicine.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Contributors<br />

Hamidreza Montazeri Aliabadi<br />

Department of Pharmacy and Pharmaceutical<br />

Sciences<br />

University of Alberta<br />

Edmonton, Alberta, Canada<br />

Christine Allen<br />

Department of Pharmaceutical Sciences<br />

and<br />

Department of Chemical Engineering and<br />

Applied Chemistry<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Ashootosh V. Ambade<br />

Department of Chemistry<br />

University of Massachusetts at Amherst<br />

Amherst, Massachusetts<br />

Mansoor M. Amiji<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Joseph M. Backer<br />

SibTech, Inc.<br />

Newington, Connecticut<br />

Marina V. Backer<br />

SibTech, Inc.<br />

Newington, Connecticut<br />

You Han Bae<br />

Department of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

Lajos P. Balogh<br />

Department of Radiation Medicine<br />

and<br />

Department of Cellular Stress and<br />

Cancer Biology<br />

Roswell Park Cancer Institute<br />

Buffalo, New York<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Rolf F. Barth<br />

Department of Pathology<br />

The Ohio State University<br />

Columbus, Ohio<br />

Reina Bendayan<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Tania Betancourt<br />

The University of Texas at Austin<br />

Austin, Texas<br />

D. Bhadra<br />

Department of Pharmaceutical Sciences<br />

Dr. H.S. Gour University<br />

Sagar, India<br />

S. Bhadra<br />

Department of Pharmaceutical Sciences<br />

Dr. H.S. Gour University<br />

Sagar, India<br />

Sangeeta N. Bhatia<br />

Harvard–MIT Division of Health Sciences<br />

and Technology<br />

Massachusetts Institute of Technology<br />

Cambridge, Massachusetts<br />

Sarathi V. Boddapati<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Lisa Brannon-Peppas<br />

The University of Texas at Austin<br />

Austin, Texas<br />

Mark Butler<br />

Department of Chemical Engineering<br />

and Applied Chemistry<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

John C. Byrd<br />

Division of Hematology–Oncology<br />

NCI Comprehensive Cancer Center<br />

The Ohio State University<br />

Columbus, Ohio


Robert B. Campbell<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Shing-Ming Cheng<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Gerard G. M. D’Souza<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Marina A. Dobrovolskaia<br />

Nanotechnology Characterization Laboratory<br />

NCI Frederick<br />

Frederick, Maryland<br />

Amber Doiron<br />

The University of Texas at Austin<br />

Austin, Texas<br />

P. K. Dubey<br />

Department of Pharmaceutical Sciences<br />

Dr. H. S. Gold University<br />

Sagar, India<br />

Omid C. Farokhzad<br />

Department of Anesthesiology, Perioperative<br />

and Pain Medicine<br />

Brigham and Women’s Hospital and Harvard<br />

Medical School<br />

Boston, Massachusetts<br />

Mauro Ferrari<br />

The Brown Foundation Institute of<br />

Molecular Medicine<br />

M. D. Anderson Cancer Center<br />

and<br />

Alliance for Nanohealth<br />

The University of Texas<br />

Houston, Texas<br />

Alberto A. Gabizon<br />

Oncology Institute<br />

Shaare Zedek Medical Center and<br />

Hebrew University<br />

Jerusalem, Israel<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Jinming Gao<br />

Simmons Comprehensive Cancer Center<br />

University of Texas Southwestern<br />

Medical Center<br />

Dallas, Texas<br />

Hamidreza Ghandehari<br />

Department of Pharmaceutical Sciences<br />

University of Maryland<br />

Baltimore, Maryland<br />

Todd J. Harris<br />

Harvard–MIT Division of Health Sciences<br />

and Technology<br />

Massachusetts Institute of Technology<br />

Cambridge, Massachusetts<br />

Mitsuru Hashida<br />

Department of Drug Delivery Research<br />

Graduate School of Pharmaceutical Sciences<br />

Kyoto University<br />

Sakyo-ku, Kyoto, Japan<br />

Yuriko Higuchi<br />

Department of Drug Delivery Research<br />

Graduate School of Pharmaceutical Sciences<br />

Kyoto University<br />

Sakyo-ku, Kyoto, Japan<br />

Kang Moo Huh<br />

Department of Polymer Science<br />

and Engineering<br />

Chungnam National University<br />

Yuseong-Gu, Daejeon, South Korea<br />

Edward F. Jackson<br />

Division of Diagnostic Imaging<br />

The University of Texas M. D. Anderson<br />

Cancer Center<br />

Houston, Texas<br />

N. K. Jain<br />

Department of Pharmaceutical Sciences<br />

Dr. H.S. Gour University<br />

Sagar, India<br />

Eun-Kee Jeong<br />

Department of Radiology<br />

University of Utah<br />

Salt Lake City, Utah


Ji Hoon Jeong<br />

Department of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

Sangyong Jon<br />

Department of Life Science<br />

Gwangju Institute of Science and Technology<br />

Gwangju, South Korea<br />

Shigeru Kawakami<br />

Department of Drug Delivery Research<br />

Graduate School of Pharmaceutical Sciences<br />

Kyoto University<br />

Sakyo-ku, Kyoto, Japan<br />

Mohamed K. Khan<br />

Department of Radiation Medicine<br />

and<br />

Department of Cellular Stress and<br />

Cancer Biology<br />

Roswell Park Cancer Institute<br />

Buffalo, New York<br />

Chalermchai Khemtong<br />

Simmons Comprehensive Cancer Center<br />

University of Texas Southwestern<br />

Medical Center<br />

Dallas, Texas<br />

Sun Hwa Kim<br />

Department of Biological Sciences<br />

Korea Advanced Institute of Science<br />

and Technology<br />

Daejeon, South Korea<br />

Tae-il Kim<br />

School of Chemistry and Molecular<br />

Engineering<br />

Seoul National University<br />

Seoul, South Korea<br />

Sushma Kommareddy<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Vinod Labhasetwar<br />

Departments of Pharmaceutical Sciences and<br />

Biochemistry and Molecular Biology<br />

University of Nebraska Medical Center<br />

Omaha, Nebraska<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Andras G. Lacko<br />

Departments of Molecular Biology and<br />

Immunology<br />

University of North Texas Health Science<br />

Center<br />

Fort Worth, Texas<br />

Robert Langer<br />

Department of Chemical Engineering<br />

Massachusetts Institute of Technology<br />

Cambridge, Massachusetts<br />

Afsaneh Lavasanifar<br />

Department of Pharmacy and Pharmaceutical<br />

Sciences<br />

University of Alberta<br />

Edmonton, Alberta, Canada<br />

Eun Seong Lee<br />

Amorepacific Corporation/R&D Center<br />

Giheung-gu Yongin-si Gyeonggi-do<br />

South Korea<br />

Helen Lee<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Robert J. Lee<br />

NCI Comprehensive Cancer Center<br />

and<br />

NSF Nanoscale Science and<br />

Engineering Center<br />

The Ohio State University<br />

Columbus, Ohio<br />

Sang Cheon Lee<br />

Nanomaterials Application Division<br />

Korea Institute of Ceramic Engineering and<br />

Technology<br />

Guemcheon-gu, Seoul, South Korea<br />

Chun Li<br />

M. D. Anderson<br />

Cancer Center<br />

The University of Texas<br />

Houston, Texas<br />

Yongqiang Li<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada


Bruce R. Line<br />

Department of Radiology, and Greenebaum<br />

Cancer Center<br />

University of Maryland<br />

Baltimore, Maryland<br />

Jubo Liu<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Zheng-Rong Lu<br />

Departments of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

S. Mahor<br />

Department of Pharmaceutical Sciences<br />

Dr. H. S. Gour University<br />

Sagar, India<br />

Walter J. McConathy<br />

Department of Internal Medicine<br />

University of North Texas Health Science<br />

Center<br />

Fort Worth, Texas<br />

Scott E. McNeil<br />

Nanotechnology Characterization Laboratory<br />

NCI Frederick<br />

Frederick, Maryland<br />

T. J. Miller<br />

Center for Drug Evaluation and Research<br />

Food and Drug Administration<br />

Silver Spring, Maryland<br />

Amitava Mitra<br />

Department of Pharmaceutical Sciences and<br />

Center for Nanomedicine and<br />

Cellular Delivery<br />

University of Maryland<br />

Baltimore, Maryland<br />

S. M. Moghimi<br />

Molecular Targeting and Polymer<br />

Toxicology <strong>Group</strong><br />

School of Pharmacy<br />

University of Brighton<br />

Brighton, United Kingdom<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Randall J. Mrsny<br />

Center for Drug Delivery/Biology<br />

Welsh School of Pharmacy<br />

Cardiff University<br />

Cardiff, United Kingdom<br />

R. S. R. Murthy<br />

Pharmacy Department<br />

The M.S. University of Baroda<br />

Vadodara, India<br />

Natarajan Muthusamy<br />

NCI Comprehensive Cancer Center<br />

The Ohio State University<br />

Columbus, Ohio<br />

Anjan Nan<br />

Department of Pharmaceutical Sciences and<br />

Center for Nanomedicine and<br />

Cellular Delivery<br />

University of Maryland<br />

Baltimore, Maryland<br />

Maya Nair<br />

Departments of Molecular Biology and<br />

Immunology<br />

University of North Texas Health<br />

Science Center<br />

Fort Worth, Texas<br />

Norased Nasongkla<br />

Simmons Comprehensive Cancer Center<br />

University of Texas Southwestern<br />

Medical Center<br />

Dallas, Texas<br />

David Needham<br />

Department of Mechanical Engineering<br />

and Materials Science<br />

Duke University<br />

Durham, North Carolina<br />

Tooru Ooya<br />

Department of Materials Science<br />

Japan Advanced Institute of Science<br />

and Technology<br />

Tatsunokuchi, Ishikawa, Japan


Jong-Sang Park<br />

Department of Chemistry and Molecular<br />

Engineering<br />

Seoul National University<br />

Seoul, South Korea<br />

Kinam Park<br />

Departments of Pharmaceutics and<br />

Biomedical Engineering<br />

Purdue University<br />

West Lafayette, Indiana<br />

Tae Gwan Park<br />

Department of Biological Sciences<br />

Korea Advanced Institute of Science and<br />

Technology<br />

Daejeon, South Korea<br />

Anil K. Patri<br />

Nanotechnology Characterization Laboratory<br />

NCI Frederick<br />

Frederick, Maryland<br />

Ana Ponce<br />

Department of Biomedical Engineering<br />

Duke University<br />

Durham, North Carolina<br />

Natalya Rapoport<br />

Department of Bioengineering<br />

University of Utah<br />

Salt Lake City, Utah<br />

Mike Andrew Rauth<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

L. Harivardan Reddy<br />

Pharmacy Department<br />

The M.S. University of Baroda<br />

Vadodara, India<br />

Fredika M. Robertson<br />

College of Medicine and Comprehensive<br />

Cancer Center<br />

The Ohio State University<br />

Columbus, Ohio<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

N. Sadrieh<br />

Center for Drug Evaluation and Research<br />

Food and Drug Administration<br />

Silver Spring, Maryland<br />

Sanjeeb K. Sahoo<br />

Department of Pharmaceutical Sciences<br />

Nebraska Medical Center<br />

Omaha, Nebraska<br />

and<br />

Institute of Life Sciences<br />

Bhubaneswar, India<br />

Elamprakash N. Savariar<br />

Department of Chemistry<br />

University of Massachusetts at Amherst<br />

Amherst, Massachusetts<br />

Dinesh B. Shenoy<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts and Novavax, Inc.<br />

Malvern, Pennsylvania<br />

Patrick Lim Soo<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Stephan T. Stern<br />

Nanotechnology Characterization<br />

Laboratory<br />

NCI Frederick<br />

Frederick, Maryland<br />

S. (Thai) Thayumanavan<br />

Department of Chemistry<br />

University of Massachusetts at Amherst<br />

Amherst, Massachusetts<br />

Sandip B. Tiwari<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Werner Tjarks<br />

College of Pharmacy<br />

The Ohio State University<br />

Columbus, Ohio


Vladimir P. Torchilin<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

Anagha Vaidya<br />

Departments of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

Geoffrey von Maltzahn<br />

Harvard–MIT Division of Health Sciences<br />

and Technology<br />

Massachusetts Institute of Technology<br />

Cambridge, Massachusetts<br />

S. P. Vyas<br />

Department of Pharmaceutical Sciences<br />

Dr. H. S. Gour University<br />

Sagar, India<br />

Sidney Wallace<br />

Division of Diagnostic Imaging<br />

The University of Texas M. D. Anderson<br />

Cancer Center<br />

Houston, Texas<br />

Yanli Wang<br />

Departments of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

Volkmar Weissig<br />

Department of Pharmaceutical Sciences<br />

Northeastern University<br />

Boston, Massachusetts<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Ho Lun Wong<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Gong Wu<br />

Department of Pathology<br />

The Ohio State University<br />

Columbus, Ohio<br />

Xiao Yu Wu<br />

Department of Pharmaceutical Sciences<br />

University of Toronto<br />

Toronto, Ontario, Canada<br />

Xiao-Bing Xiong<br />

Faculty of Pharmacy and Pharmaceutical<br />

Sciences<br />

University of Alberta<br />

Edmonton, Alberta, Canada<br />

Weilian Yang<br />

Department of Pathology<br />

The Ohio State University<br />

Columbus, Ohio<br />

Furong Ye<br />

Departments of Pharmaceutics and<br />

Pharmaceutical Chemistry<br />

University of Utah<br />

Salt Lake City, Utah<br />

Guodong Zhang<br />

M. D. Anderson<br />

Cancer Center<br />

The University of Texas<br />

Houston, Texas<br />

Xiaobin B. Zhao<br />

NCI Comprehensive Cancer Center<br />

The Ohio State University<br />

Columbus, Ohio<br />

.


Table of Contents<br />

Section 1<br />

Nanotechnology and Cancer. .................................................1<br />

Chapter 1 Introduction and Rationale for Nanotechnology<br />

in Cancer Therapy . ..............................................3<br />

Fredika M. Robertson and Mauro Ferrari<br />

Chapter 2 Passive Targeting of Solid Tumors: Pathophysiological<br />

Principles and Physicochemical Aspects of<br />

Delivery Systems . . .............................................11<br />

S. M. Moghimi<br />

Chapter 3 Active Targeting Strategies in Cancer with a Focus on<br />

Potential Nanotechnology Applications . ..............................19<br />

Randall J. Mrsny<br />

Chapter 4 Pharmacokinetics of Nanocarrier-Mediated<br />

Drug and Gene Delivery . . . ......................................43<br />

Yuriko Higuchi, Shigeru Kawakami, and Mitsuru Hashida<br />

Chapter 5 Multifunctional Nanoparticles for Cancer Therapy. . . . ...................59<br />

Todd J. Harris, Geoffrey von Maltzahn, and Sangeeta N. Bhatia<br />

Chapter 6 Neutron Capture Therapy of Cancer: Nanoparticles<br />

and High Molecular Weight Boron Delivery Agents. . . ...................77<br />

Gong Wu, Rolf F. Barth, Weilian Yang, Robert J. Lee, Werner Tjarks, Marina V. Backer, and<br />

Joseph M. Backer<br />

Chapter 7 Preclinical Characterization of Engineered Nanoparticles<br />

Intended for Cancer Therapeutics . .................................105<br />

Anil K. Patri, Marina A. Dobrovolskaia, Stephan T. Stern, and Scott E. McNeil<br />

Chapter 8 Nanotechnology: Regulatory Perspective for Drug<br />

Development in Cancer Therapeutics . . .............................139<br />

N. Sadrieh and T. J. Miller<br />

Section 2<br />

Polymer Conjugates. . ....................................................157<br />

Chapter 9 Polymeric Conjugates for Angiogenesis Targeted Tumor<br />

Imaging and Therapy . . . . . . .....................................159<br />

Amitava Mitra, Anjan Nan, Bruce R. Line, and Hamidreza Ghandehari<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Chapter 10 Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and<br />

Diagnostics . . . . . ............................................185<br />

Guodong Zhang, Edward F. Jackson, Sidney Wallace, and Chun Li<br />

Chapter 11 Noninvasive Visualization of In Vivo Drug Delivery<br />

of Paramagnetic Polymer Conjugates with MRI. . .....................201<br />

Zheng-Rong Lu, Yanli Wang, Furong Ye, Anagha Vaidya, and Eun-Kee Jeong<br />

Section 3<br />

Polymeric Nanoparticles . . . . . . ............................................213<br />

Chapter 12 Polymeric Nanoparticles for Tumor-Targeted Drug<br />

Delivery. ...................................................215<br />

Tania Betancourt, Amber Doiron, and Lisa Brannon-Peppas<br />

Chapter 13 Long-Circulating Polymeric Nanoparticles for Drug and<br />

Gene Delivery to Tumors . . . ....................................231<br />

Sushma Kommareddy, Dinesh B. Shenoy, and Mansoor M. Amiji<br />

Chapter 14 Biodegradable PLGA/PLA Nanoparticles for Anti-cancer<br />

Therapy . ...................................................243<br />

Sanjeeb K. Sahoo and Vinod Labhasetwar<br />

Chapter 15 Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery<br />

of Anti-Cancer Drugs . . ........................................251<br />

R. S. R. Murthy and L. Harivardhan Reddy<br />

Chapter 16 Aptamers and Cancer Nanotechnology .............................289<br />

Omid C. Farokhzad, Sangyong Jon, and Robert Langer<br />

Section 4<br />

Polymeric Micelles . . . ...................................................315<br />

Chapter 17 Polymeric Micelles for Formulation of Anti-Cancer Drugs. . . . . ..........317<br />

Helen Lee, Patrick Lim Soo, Jubo Liu, Mark Butler, and Christine Allen<br />

Chapter 18 PEO-Modified Poly(L-Amino Acid) Micelles for<br />

Drug Delivery . . . ............................................357<br />

Xiao-Bing Xiong, Hamidreza Montazeri Aliabadi, and Afsaneh Lavasanifar<br />

Chapter 19 Hydrotropic Polymer Micelles for Cancer Therapeutics .................385<br />

Sang Cheon Lee, Kang Moo Huh, Tooru Ooya, and Kinam Park<br />

Chapter 20 Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer<br />

Drugs with Polymeric Lipid-Core Immunomicelles . . . .................409<br />

Vladimir P. Torchilin<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Chapter 21 Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated<br />

Drugs and Ultrasound. . . . . .....................................421<br />

Natalya Rapoport<br />

Chapter 22 Polymeric Micelles Targeting Tumor pH. . . . . . ......................443<br />

Eun Seong Lee and You Han Bae<br />

Chapter 23 cRGD-Encoded, MRI-Visible Polymeric Micelles<br />

for Tumor-Targeted Drug Delivery . . . .............................465<br />

Jinming Gao, Norased Nasongkla, and Chalermchai Khemtong<br />

Chapter 24 Targeted Antisense Oligonucleotide Micellar<br />

Delivery Systems . ............................................477<br />

Ji Hoon Jeong, Sun Hwa Kim, and Tae Gwan Park<br />

Section 5<br />

Dendritic Nanocarriers. ...................................................487<br />

Chapter 25 Dendrimers as Drug and Gene Delivery Systems ......................489<br />

Tae-il Kim and Jong-Sang Park<br />

Chapter 26 Dendritic Nanostructures for Cancer Therapy . . ......................509<br />

Ashootosh V. Ambade, Elamprakash N. Savariar, and S. (Thai) Thayumanavan<br />

Chapter 27 PEGylated Dendritic Nanoparticulate Carriers of<br />

Anti-Cancer Drugs ............................................523<br />

D. Bhadra, S. Bhadra, and N. K. Jain<br />

Chapter 28 Dendrimer Nanocomposites for Cancer Therapy ......................551<br />

Lajos P. Balogh and Mohamed K. Khan<br />

Section 6<br />

Liposomes . . . . ........................................................593<br />

Chapter 29 Applications of Liposomal Drug Delivery Systems<br />

to Cancer Therapy ............................................595<br />

Alberto A. Gabizon<br />

Chapter 30 Positively-Charged Liposomes for Targeting<br />

Tumor Vasculature. ...........................................613<br />

Robert B. Campbell<br />

Chapter 31 Cell Penetrating Peptide (CPP)–Modified Liposomal<br />

Nanocarriers for Intracellular Drug and Gene Delivery . . . ..............629<br />

Vladimir P. Torchilin<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Chapter 32 RGD-Modified Liposomes for Tumor Targeting . .....................643<br />

P. K. Dubey, S. Mahor, and S. P. Vyas<br />

Chapter 33 Folate Receptor-Targeted Liposomes for<br />

Cancer Therapy . . ............................................663<br />

Xiaobin B. Zhao, Natarajan Muthusamy, John C. Byrd, and Robert J. Lee<br />

Chapter 34 Nanoscale Drug Delivery Vehicles for Solid Tumors:<br />

A New Paradigm for Localized Drug Delivery<br />

Using Temperature Sensitive Liposomes . . . .........................677<br />

David Needham and Ana Ponce<br />

Section 7<br />

Other Lipid Nanostructures . . . . ............................................721<br />

Chapter 35 Nanoemulsion Formulations for Tumor-Targeted Delivery . . . . . ..........723<br />

Sandip B. Tiwari and Mansoor M. Amiji<br />

Chapter 36 Solid Lipid Nanoparticles for Antitumor Drug Delivery .................741<br />

Ho Lun Wong, Yongqiang Li, Reina Bendayan, Mike Andrew Rauth, and Xiao Yu Wu<br />

Chapter 37 Lipoprotein Nanoparticles as Delivery Vehicles<br />

for Anti-Cancer Agents ........................................777<br />

Andras G. Lacko, Maya Nair, and Walter J. McConathy<br />

Chapter 38 DQAsomes as Mitochondria-Targeted Nanocarriers<br />

for Anti-Cancer Drugs . ........................................787<br />

Shing-Ming Cheng, Sarathi V. Boddapati, Gerard G. M. D’Souza, and Volkmar Weissig<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


1<br />

CONTENT<br />

Introduction and Rationale<br />

for Nanotechnology<br />

in Cancer Therapy<br />

Fredika M. Robertson and Mauro Ferrari<br />

References ........................................................................................................................................ 8<br />

Although there have been significant advances in defining the fundamentals of cancer biology over<br />

the past 25 years, this has not translated into similar clinical advances in cancer therapeutics. One<br />

area that holds great promise for making such advances is the area defined as cancer nanotechnology,<br />

which involves the intersection of a variety of disciplines, including engineering, materials<br />

science, chemistry, and physics with cancer biology. This multidisciplinary convergence has<br />

resulted in the creation of devices and/or materials that are themselves or have essential components<br />

in the 1–1000-nm range for at least one dimension and holds the possibility of rapidly<br />

advancing the state of cancer therapeutics and tumor imaging.<br />

This newly developing area of “nanohealth” may ultimately allow detection of human tumors at<br />

the very earliest stages, regardless of the location of the primary tumor and/or metastases, and may<br />

provide approaches to more effectively destroy tumors as well as their associated vascular supplies<br />

with fewer adverse side effects. The chapters within this book describe the most recent cutting-edge<br />

approaches in nanotechnology that are focused on overcoming the multiple barriers that have, in the<br />

past, blocked the successful treatment and ultimate eradication of human cancers.<br />

The majority of nanotechnology-based devices useful for cancer therapeutics have been defined<br />

as nanovectors, which are injectable nanoscale delivery systems. 1,2 Nanovectors offer the promise<br />

of providing breakthrough solutions to the problems of optimizing efficacy of therapeutic agents<br />

while simultaneously diminishing the deleterious side-effects that commonly accompany the use of<br />

both single chemotherapeutic agents as well as multimodality therapeutic regimens.<br />

In general, nanovectors are comprised of at least three constituents, which include a core<br />

material, a therapeutic and/or imaging “payload,” and a biological surface modification, which<br />

aids in both appropriate biodistribution and selective localization of the nanovector and its cytotoxic<br />

and/or imaging agent. Although the first type of molecules used to enhance the selective<br />

localization and delivery of nanovectors were antibodies, more sophisticated recognition systems<br />

have been devised as a result of our expanding knowledge base in cancer biology.<br />

For example, while the biological and molecular characteristics of human tumors of different<br />

origins continue to be defined and exploited for cancer therapeutics and tumor imaging, the significance<br />

of blood vessels that develop around actively growing tumors as potential therapeutic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

3


4<br />

Nanotechnology for Cancer Therapy<br />

targets has only been widely recognized and found clinical utility within the past decade. 3,4 The<br />

process of tumor-associated angiogenesis is now known to be an essential component of tumor<br />

expansion and metastasis. This realization and the accompanied potential for development of<br />

successful cancer therapeutics, as well as for tumor imaging strategies based on tumor-associated<br />

vasculature, has opened new doors for the application of nanotechnology in cancer therapeutics.<br />

Additionally, the molecules that drive the process of tumor angiogenesis may provide a means to<br />

gauge the timing and extent of individual patient responses to cancer treatment, which can also be<br />

monitored using nanovector approaches. The importance of tumor neovasculature in cancer nanotechnology<br />

is highlighted throughout the chapters in Nanotechnology for Cancer Therapy.<br />

One of the most important characteristics of nanovectors is their ability to be functionalized to<br />

overcome barriers that block access of agents used for treatment of tumors and for imaging of<br />

tumors and their associated vasculature. These biological barriers are numerous and complex. One<br />

such barrier is the blood–brain barrier, which prevents access to brain malignancies, compounding<br />

the difficulties in their successful treatment. One example of the potential utility of nanovectors in<br />

overcoming this biobarrier, which is critical to treatment of malignant brain tumors, is the use of<br />

nanoparticles in combination with boron neutron capture therapy (BCNT). The current approaches<br />

used for BCNT combining nanoparticles and the potential for this nanotechnology-based approach<br />

to treatment of brain tumors is described in detail in a chapter in Nanotechnology for Cancer<br />

Therapy. Additional biobarriers that must be overcome include epithelial–endothelial cell barriers,<br />

the barriers presented <strong>by</strong> the markedly tortuous structures that are characteristic of angiogenic<br />

vasculature associated with tumors, as well as the barrier set up <strong>by</strong> the rapid uptake of nanovectors<br />

<strong>by</strong> resident macrophages within the reticuloendothelial system (RES) which may prevent nanovectors<br />

from reaching their targeted location.<br />

To achieve breakthrough advances in cancer therapeutics, there are two related and essential<br />

components which must be addressed. The first issue in successful use of nanovectors is recognition<br />

of the tumor and the second is the ability of the nanovector to reach the site of the tumor and<br />

associated blood vessels. The goal is to preferentially achieve high concentrations of a specific<br />

chemotherapeutic agent, a tumor imaging agent, and/or gene therapies at the site(s) of tumors and<br />

associated vasculature. In addition, nanovectors must be able to deliver an active agent to achieve<br />

effective anti-tumor treatment, or tumor imaging, which is essential for tumor diagnosis and for<br />

monitoring the extent and timing of an individual patient’s response to anti-tumor therapy.<br />

The first nanotechnology-based approach to be used as a means of delivering cancer<br />

chemotherapy was liposomes, which is a type of nanovector made of lipids surrounding a water<br />

core. Liposomes are the simplest form of nanovector and their utility is based on the significant<br />

difference in endothelial structures—defined as fenestrations—between normal vasculature and<br />

tumor-associated vessels. The increase in fenestrations in tumor neovasculature allows the preferential<br />

concentration of liposome-encapsulated anti-tumor agent in close proximity to the local<br />

tumor site, a phenomenon defined as enhanced penetration and retention (EPR), which is<br />

considered to be a characteristic of passive targeting of tumors. Both the passive targeting of<br />

solid tumors as well as approaches for active targeting of nanovectors such as liposomes are<br />

described in the following chapters.<br />

The first studies to demonstrate the proof of principle that liposomes could be used as a<br />

nanovector for effective delivery of an active cancer therapeutic was based on the encapsulation<br />

of the anthracycline antibiotic doxorubicin, which is an effective and commonly used first line<br />

chemotherapeutic agent for both leukemia and solid tumors. Although use of doxorubicin is<br />

associated with dose limiting cardiotoxicity, when anionic liposomes were used to encapsulate<br />

doxorubicin, preclinical studies found that the anti-tumor action of doxorubicin was significantly<br />

enhanced and the cardiotoxicity was significantly diminished. 5,6 Liposome encapsulated doxorubicin<br />

has now been clearly demonstrated safe and efficacious clinically in such malignancies as<br />

breast and ovarian cancer, 7–9 allowing for greater amounts of this agent to be administered than the<br />

recommended lifetime cumulative dose of free doxorubicin, which was 450–550 mg/m 2 .<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Introduction and Rationale for Nanotechnology in Cancer Therapy 5<br />

Importantly, liposomes have been shown to be an effective means of delivering a diverse group of<br />

anti-tumor agents such as doxorubicin, as well as the poorly soluble drug paclitaxel. 10<br />

As evidenced <strong>by</strong> the chapters in this volume, interest in improving the use of liposomes for<br />

enhancing the selective localization and delivery of cancer therapeutics to tumors and tumorassociated<br />

blood vessels remains an active area of investigation. Because liposomes can be<br />

modified with respect to composition, charge, and external coating, liposomes remain a versatile<br />

nanotechnology platform, historically serving as the prototypic nanovector.<br />

An approach that was first used with liposome nanovectors that has found utility with other<br />

types of nanovectors is “PEGylation,” a modification of liposome surface characteristics using<br />

poly(ethylene glycol) (PEG), resulting in what has been defined as “sterically stabilized liposomes.”<br />

This PEG modification of liposomes provides protection against uptake <strong>by</strong> resident<br />

macrophages within the RES biobarrier, thus increasing the circulation time of liposome-encapsulated<br />

anti-tumor agent, resulting in significantly increased therapeutic efficacy. 11<br />

This approach to overcoming rapid uptake and destruction <strong>by</strong> resident macrophages within the<br />

RES <strong>by</strong> PEGylation has been used alone or in tandem with other modifications of liposomes to aid<br />

in avoiding or overcoming biobarriers and to more selectively localize nanovectors. For example,<br />

based on the recognition that a wide variety of solid tumors as well as hematologic malignancies<br />

express amplified levels of the folate receptor on their cell surface, liposome nanovectors have been<br />

developed to target folate receptors. 12,13 Using the strategy of biomarker-based targeting, folate<br />

receptor targeted liposomes have been shown to not only overcome multidrug resistance associated<br />

with P-glycoprotein that often occurs in human tumors, but also have proven to be antiangiogenic. 13<br />

Another means of more selectively localizing liposome-encapsulated anti-tumor agents to tumorassociated<br />

vasculature is to add a cationic charge, 10,14 which has been shown to double the<br />

accumulation of anti-tumor drug encapsulated into liposomes selectively within vessels<br />

surrounding tumors. Other approaches to biomolecular targeting have included the use of specific<br />

proteins such as vasoactive intestinal peptide (VIP) 15 to target gastric tumors as well as signature<br />

RGD amino acid sequences expressed on the surface of endothelial cells that are associated with<br />

expression of a vb 3 integrin on angiogenic endothelium. 16<br />

In addition to serving as cancer drug delivery systems, liposomes have been shown to be an<br />

effective means for delivery of other agents such as genes and antisense oligonucleotides, and<br />

would allow access of such entities as small interfering RNA. 13,17 As an example, peptides such as<br />

cell penetrating peptides (CPPs) have been used to modify liposomes, allowing them to be used as a<br />

means for intracellular drug delivery and delivery of proteins such as antibodies and genes, as well<br />

as providing a window for cellular imaging. 17<br />

In addition to liposomes as an example of the prototype of nanovectors, there are many others<br />

now available in the nanotechnology toolbox that are being investigated for use in cancer therapeutics<br />

and tumor imaging. Polymer-based nanovectors are a specific area of interest in cancer<br />

therapeutics and a number of polymer-based nanovector systems are described in the following<br />

chapters. For the purposes of this volume, polymer-based nanovectors have been subdivided into<br />

several categories, including polymer conjugates, polymeric nanoparticles, and polymeric micelles.<br />

Polymeric conjugates that have been investigated for use as nanovector systems for<br />

delivery of anti-tumor agents include the water soluble biocompatible polymer N-(2hydroxypropyl)methacrylamide<br />

(HPMA). This polymeric conjugate has been targeted based on<br />

overexpression of hyaluronan (HA) receptors that are present on cancer cells. HPMA–HA polymeric<br />

conjugate drug delivery nanovector systems have been developed to carry a doxorubicin “payload”<br />

and have been shown to have selectivity for endocytosis of the targeted polymeric nanovector to<br />

breast, ovarian, and colon tumor cells, compared to an HPMA polymer with doxorubicin but lacking<br />

the HA targeting conjugate. 18,19 Copolymer–peptide conjugates have also been developed as nanovectors<br />

using HPMA in combination with RGD targeting peptides, 20 which molecularly target<br />

radiotherapeutic agents to tumor-associated vasculature, resulting in both antiangiogenic as well<br />

as anti-tumor activity. In addition to serving as nanovectors for delivery of chemotherapy and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


6<br />

Nanotechnology for Cancer Therapy<br />

radiotherapy, copolymer nanovectors show promise as agents to deliver biodegradable blood-pool<br />

contrast agents for use with tumor imaging methodologies. Other copolymers that have been used<br />

with gadolinium chelates for improvements in MRI based contrast agents include poly(L-glutamic<br />

acid) 21 as well as (Gd-DTPA)-cystine copolymers (GDCP) and (Gd-DTPA)-cystine diethyl ester<br />

copolymers (GDCEP). 22,23 As with liposome nanovectors, polymer conjugates can be targeted to a<br />

specific site, can carry a multitude of agents useful for cancer chemotherapeutics, radiotherapy, or<br />

tumor imaging, and their surfaces can be modified using different biological surface modifiers to<br />

enhance selectivity of localization of the nanovector.<br />

Studies focusing on polymeric nanoparticles as nanovectors have included the use of poly(lacticco-glycolic<br />

acid) (PLGA) and polylactide (PLA) that can be selectively targeted against such surface<br />

receptors as the transferrin receptor on breast tumor cells, as well as can be optimized to avoid<br />

biobarriers such as the RES. In addition, polymeric nanoparticles can be altered to enhance the<br />

intracellular drug concentration once delivered. Polymeric nanoparticles are also mutable for<br />

enhancement of retention at the local tumor site and to combine anti-tumor agents with antiangiogenesis<br />

strategies. 24–26 In addition to delivery of anti-tumor agents that are antiangiogenic,<br />

nanovectors such as PLGA and PLA polymers as well as PEGylated gelatin nanoparticles have<br />

been developed. In the case of PEGylated gelatin nanovectors, these have been improved <strong>by</strong> thiolation<br />

to optimize their time in circulation and to enhance delivery of payloads such as non-viral DNA<br />

for gene therapy. 25,27–29<br />

Relatively new polymeric nanovectors recently shown to have potential utility in cancer nanotechnology<br />

are bioconjugates composed of PLGA polymers and nucleic acid ligands defined as<br />

aptamers. 30 Because these nanovectors can be PEGylated and further modified with selectively<br />

targeted approaches directed against such proteins as prostate specific antigen (PSA) which are<br />

present in abundance on the surface of prostatic tumor cells, these nanoparticle–aptamer bioconjugates<br />

represent a new class of polymeric nanoparticles with promise as multifunctional<br />

nanovectors. Polymeric aptamer bioconjugates are currently being evaluated for their optimization<br />

for use in a large number of systems that may lead to in vivo studies in the near future. 31<br />

Another type of nanovector discussed in Nanotechnology for Cancer Therapy includes polymeric<br />

micelles, which can be defined as self-assembling nanosized colloidal particles with a<br />

hydrophobic core and hydrophilic shell. 32 Polymeric micelles have been commonly used in the<br />

field of drug delivery for over two decades. The advantages of these nanovectors include the ability<br />

to encapsulate agents with poor aqueous solubility profiles such as has been observed with the antitumor<br />

agent paclitaxel 33 as well as with the photodynamic therapy agent, meso-tetratphenylporphine<br />

(TPP). 34 Other advantages associated with the use of polymeric micelles include the ability to<br />

control both drug targeting using immunotargeting 34 and rates of drug release. 32,35 The efficacy of<br />

agents encapsulated within polymeric micelle nanovectors occurs via the passive targeting method<br />

defined as enhanced penetration and retention (EPR), which results in diminished adverse side<br />

effects usually observed in normal tissues <strong>by</strong> administration of cytotoxic cancer<br />

chemotherapeutic agents.<br />

In addition to the utility of polymeric micelles for delivery of anti-cancer agents, 32–35 these<br />

nanovectors are also described as an effective means for the combined delivery of anti-tumor agents<br />

as well as simultaneous tumor imaging using focused ultrasound imaging. This makes it possible to<br />

non-invasively penetrate tumors that are typically difficult to access, such as with ovarian cancer. 36<br />

Polymeric micelles have also found usefulness as multifunctional nanovectors that are both visible<br />

to MRI imaging modalities using RGD-based targeting of the tumor-associated vasculature<br />

combined with drug delivery. Other approaches using polymeric micelles have focused on targeting<br />

the acidic extracellular pH of the tumor microenvironment. 37 Polymeric micelles have also been<br />

evaluated for delivery of antisense oligonucleotides using folate receptor targeting strategies. 38<br />

Although polymeric micelle nanovectors have been used as strategies for drug delivery, it is clear<br />

that they continue to have promise in cancer nanotechnology.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Introduction and Rationale for Nanotechnology in Cancer Therapy 7<br />

Another class of nanovectors described in the following chapters is dendrimers, which are self<br />

assembling synthetic polymers possessing tunable nanoscale dimensions that assume conformations<br />

that are dependent upon the microenvironment. The development of polymeric micelles<br />

preceded the development of dendrimers. However, due to the characteristics of dendrimers as<br />

“perfectly branched monodispersed macromolecules,” 39 this class of nanovectors may be suitable<br />

as exquisitely sensitive carriers for defined release of anti-tumor agents, as well as very useful for<br />

targeted delivery of such molecules as contrast agents for tumor imaging.<br />

Dendritic nanocarriers have been evaluated for their capability to serve as nanovectors for<br />

delivery of drugs and for plasmid DNA, suggesting that dendrimer nanovectors may be useful<br />

for gene therapy. 40 Like liposomes, dendrimers can also be modified <strong>by</strong> PEGylation 41 to increase<br />

their bioavailability and half-life in the circulation. Dendrimers have been shown to be suitable for<br />

enhancing the solubility of drugs, and to have suitable profiles for controlled delivery as well as<br />

selective delivery of drugs such as flurbiprofen that are active within a local site of inflammation. 42<br />

Although dendrimers that have been evaluated for potential utility in cancer therapeutics have<br />

primarily been composed of polyamidoamine (PAMAM), there are also gold/dendrimer nanocomposites<br />

43 and silver/dendrimer nanocomposite materials 44 that have been developed and are<br />

currently being evaluated for their comparative chemical and biological characteristics with<br />

potential utility as biomarkers, as well as for use in cancer therapeutics, tumor imaging, and<br />

radiotherapy.<br />

The varied nature of nanovectors covered in Nanotechnology for Cancer Therapy is illustrated<br />

<strong>by</strong> such novel classes of nanocarriers as DQAsomes, which are mitochondrial-targeted nanovectors<br />

capable of delivery of anti-cancer drugs as well as carriers of gene therapy directed against this<br />

organelle. 44,45 The development of this nanovector delivery approach is based on the relatively<br />

recent realization that a number of genetic disorders associated with defects in oxidative metabolism<br />

are associated with genetic defects in mitochondrial DNA, with a significant lack of available<br />

therapies. While the majority of nanovectors target tumors and/or tumor-associated vasculature, the<br />

DQAsomes are composed of derivatives of the self-assembling mitochondriotropic bola-amphiphile<br />

dequalinium chloride, which forms cationic vesicles which can bind and transport DNA to<br />

mitochondria in living mammalian cells. 44<br />

Another approach using selective targeting of more classically used characteristics of cancer<br />

cells includes the development of lipoprotein nanoparticles as delivery vehicles for anti-cancer<br />

therapy. 46 These high density lipoprotein (rHDL) nanoparticles are composed of phosphatidylcholine,<br />

apolipoprotein A-1, cholesterol, and cholesteryl esters with encapsulated paclitaxel.<br />

They were shown to be useful as a nanovector for anti-tumor drug delivery based on the ability<br />

of cancer cells to selectively acquire HDL, while the use of these nanovectors decreased the side<br />

effects normally observed with this type of anti-tumor therapy.<br />

Given the wide variety of nanovectors and nanoplatforms, Nanotechnology for Cancer Therapy<br />

cannot possibly cover each nanoscale application that has potential for use in anti-tumor therapy<br />

and tumor imaging. One nanoplatform that should be mentioned is the so-called “quantum dots”<br />

(Q-dots) that have recently received wide attention. Although cadmium and selenium crystalline<br />

nanovectors were first used over two decades ago, Q-dots have recently received attention for their<br />

utility as tunable multicolor nanoparticles that can be used for in vivo tumor imaging in animal<br />

models of cancer as well as for multiplex detection of mRNAs and proteins. Although the current<br />

composition of the Q-dots nanovectors prohibits their clinical use for cancer therapy, there is<br />

significant promise for the potential for development of biocompatible nanovectors which are<br />

tunable to wavelengths useful for tumor imaging. They may provide the ability to selectively<br />

target tumors as well as serve as multifunctional nanovector devices for cancer therapy,<br />

imaging, diagnostics, and monitoring of response to anti-tumor therapy.<br />

Although the development of nanoplatforms for cancer therapeutics is a critical component to<br />

the successful use of nanovectors for anti-tumor therapy and tumor imaging, there are other necessary<br />

components for nanoplatforms to gain clinical acceptance. One such parameter is the necessary<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


8<br />

infrastructure to characterize nanoplatforms and to ensure that the rapid translation of discoveries in<br />

this area proceeds from preclinical to clinical use. The recognition <strong>by</strong> the National Cancer Institute<br />

(NCI) that nanodevices and nanomaterials will be critical in making significant advances in<br />

imaging, diagnosis, and treatment of human tumors has lead to the development of The Nanotechnology<br />

Characterization Laboratory (NCL). This laboratory provides critical support to the NCI’s<br />

Alliance in Nanotechnology for Cancer. The role and function of the services provided <strong>by</strong> NCL is<br />

supported <strong>by</strong> the NCI, and works in concert with the National Institute of Standards and Technology<br />

(NIST) and the U.S. Food and Drug Administration (FDA). The NCL provides a valuable<br />

service to those investigators working in the area of cancer nanotechnology and performs preclinical<br />

efficacy and toxicity testing of nanoscale devices and materials. The goal of the NCL is to<br />

support the rapid acceleration of preclinical development of nanovectors to support IND application<br />

and subsequent clinical trials.<br />

The continued development and ultimate success of nanotechnology-based strategies for cancer<br />

therapeutics, imaging, and tracking patient therapeutic response will be dependent upon a number<br />

of parameters. A multidisciplinary approach to achieving success in cancer nanotechnology is<br />

absolutely necessary. While remaining grounded in our basic knowledge of the known targets<br />

that are critical for localization of multifunctional nanovectors to the tumor and surrounding<br />

tissue, researchers must be aware of new targets that have been identified as being critically<br />

involved in the response of tumors and tumor vasclulature to anti-tumor agents, and to recognize<br />

and take full advantage of known targets such as surface receptors that are amplified on tumor cells<br />

as well as recognize and incorporate newer targets such as mitochondrial DNA and HDL uptake.<br />

For example, while researchers continue to identify novel means to functionalize nanovectors to<br />

eliminate biobarriers that historically have limited access to tumors and their associated angiogenic<br />

vasculature and stroma, we will still need to produce and test multifunctional nanovectors capable<br />

of selective localization and delivery of agents that function both as cytotoxic agents as well as<br />

imaging agents.<br />

Because the size of nanodevices and nanomaterials is similar to that of naturally occurring<br />

components of cells, including surface receptors, organelles such as mitochondria, lipoproteins, and<br />

other, as yet unidentified molecules such as siRNA, microRNA, genes, and unknown proteins,<br />

nanodevices and nanomaterials can easily interface with biological molecules. The mutability of<br />

nanotechnology platforms has allowed their rapid introduction into the field of basic cancer biology<br />

and the specialty of oncology. The approaches described in the chapters within Nanotechnology for<br />

Cancer Therapy range from liposomes, which are among the first and the simplest nanovectors to<br />

be used for delivery of cytotoxic drugs, to advanced nanoscale devices for use as imaging contrast<br />

agents in conjunction with ultrasound and magnetic resonance imaging (MRI). Each of the nanotechnology<br />

approaches described in this book will be absolutely necessary to achieve significant<br />

breakthroughs in cancer therapeutics. It is truly an exciting time to be an integral part of this field as<br />

these discoveries in cancer nanotechnology are made and put into practice with the goal to use<br />

nanovectors and nanotechnology platforms to significantly impact cancer morbidity and mortality<br />

in the next decade.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Ferrari, M., Nanovector therapeutics, Curr. Opin. Chem. Biol., 9, 343–346, 2005.<br />

2. Ferrari, M., Cancer nanotechnology: Opportunities and challenges, Nat. Rev. Cancer, 5(3), 161–171,<br />

2005.<br />

3. Folkman, J., Fundamental concepts of the angiogenic process, Curr. Mol. Med., 3(7), 643–651, 2003.<br />

4. Ferrara, N. and Kerbel, R. S., Angiogenesis as a therapeutic target, Nature, 438, 967–974, 2005.<br />

5. Treat, J., Greenspan, A., Forst, D., Sanchez, J. A., Ferrans, V. J., Potkul, L. A., Woolley, P. V., and<br />

Rahman, A., Antitumor activity of liposome-encapsulated doxorubicin in advanced breast cancer:<br />

Phase II study, J. Natl. Cancer Inst., 82(21), 1706–1710, 1990.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Introduction and Rationale for Nanotechnology in Cancer Therapy 9<br />

6. Forssen, E. A. and Tokes, Z. A., Improved therapeutic benefits of doxorubicin <strong>by</strong> entrapment in<br />

anionic liposomes, Cancer Res., 43(2), 546–550, 1983.<br />

7. Straubinger, R. M., Lopez, N. G., Debs, R. J., Hong, K., and Papahadjopoulos, D., Liposome-based<br />

therapy of human ovarian cancer: Parameters determining potency of negatively charged and<br />

antibody-targeted liposomes, Cancer Res., 48(18), 5237–5245, 1988.<br />

8. Robert, N. J., Vogel, C. L., Henderson, I. C., Sparano, J. A., Moore, M. R., Silverman, P., Overmoyer,<br />

B. A. et al., The role of the liposomal anthracyclines and other systemic therapies in the management<br />

of advanced breast cancer, Semin. Oncol., 31(6 Suppl 13), 106–146, 2004.<br />

9. Ewer, M. S., Martin, F. J., and Henderson, Use of anionic liposomes for the reduction of chronic<br />

doxorubicin-induced cardiotoxicity, Proc. Natl. Acad. Sci. USA, 78(3), 1873–1877, 1981.<br />

10. Campbell, R. B., Balasubramanian, S. V., and Straubinger, R. M., Influence of cationic lipids on the<br />

stability and membrane properties of paclitaxel-containing liposomes, J. Pharm. Sci., 90(8),<br />

1091–1105, 2001.<br />

11. Gabizon, A., Shmeeda, H., Horowitz, A. T., and Zalipsky, S., Tumor cell targeting of liposomeentrapped<br />

drugs with phospholipid-anchored folic acid-PEG conjugates, Adv. Drug Deliv. Rev., 56(8),<br />

1177–1192, 2004.<br />

12. Gabizon, A., Horowitz, A. T., Goren, D., Tzemach, D., Shmeeda, H., and Zalipsky, S., In vivo fate of<br />

folate-targeted polyethylene-glycol liposomes in tumor-bearing mice, Clin. Cancer Res., 9(17),<br />

6551–6559, 2003.<br />

13. Pan, X. and Lee, R. J., Tumour-selective drug delivery via folate receptor-targeted liposomes, Expert<br />

Opin. Drug Deliv., 1(1), 7–17, 2004.<br />

14. Campbell, R. B., Fukumura, D., Brown, E. B., Mazzola, L. M., Izumi, Y., Jain, R. K., Torchilin, V. P.,<br />

and Munn, L. L., Cationic charge determines the distribution of liposomes between the vascular and<br />

extravascular compartments of tumors, Cancer Res., 62(23), 6831–6836, 2002.<br />

15. Sethi, V., Onyuksel, H., and Rubinstein, I., Liposomal vasoactive intestinal peptide, Methods<br />

Enzymol., 391, 377–395, 2005.<br />

16. Dubey, P. K., Mishra, V., Jain, S., Mahor, S., and Vyas, S. P., Liposomes modified with cyclic RGD<br />

peptide for tumor targeting, J. Drug Target., 12(5), 257–264, 2004.<br />

17. Gupta, B., Levchenko, T. S., and Torchilin, V. P., Intracellular delivery of large molecules and small<br />

particles <strong>by</strong> cell-penetrating proteins and peptides, Adv. Drug Deliv. Rev., 57(4), 637–651, 2005.<br />

18. Luo, Y. and Prestwich, G. D., Cancer-targeted polymeric drugs, Curr. Cancer Drug Targets, 2(3),<br />

209–226, 2002.<br />

19. Luo, Y., Bernshaw, N. J., Lu, Z. R., Kopecek, J., and Prestwich, G. D., Targeted delivery of doxorubicin<br />

<strong>by</strong> HPMA copolymer-hyaluronan bioconjugates, Pharm. Res., 19(4), 396–402, 2002.<br />

20. Mitra, A., Nan, A., Papadimitriou, J. C., Ghandehari, H., and Line, B. R., Polymer–peptide conjugates<br />

for angiogenesis targeted tumor radiotherapy, Nucl. Med. Biol., 33(1), 43–52, <strong>2006</strong>.<br />

21. Wen, X., Jackson, E. F., Price, R. E., Kim, E. E., Wu, Q., Wallace, S., Charnsangavej, C., Gelovani,<br />

J. G., and Li, C., Synthesis and characterization of poly(L-glutamic acid) gadolinium chelate: a new<br />

biodegradable MRI contrast agent, Bioconjug. Chem., 15(6), 1408–1415, 2004.<br />

22. Wang, X., Feng, Y., Ke, T., Schabel, M., and Lu, Z. R., Pharmacokinetics and tissue retention of<br />

(Gd-DTPA)-cystamine copolymers, a biodegradable macromolecular magnetic resonance imaging<br />

contrast agent, Pharm. Res., 22(4), 596–602, 2005.<br />

23. Zong, Y., Wang, X., Goodrich, K. C., Mohs, A. M., Parker, D. L., and Lu, Z. R., Contrast-enhanced<br />

MRI with new biodegradable macromolecular Gd(III) complexes in tumor-bearing mice, Magn.<br />

Reson. Med., 53(4), 835–842, 2005.<br />

24. Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancertherapy, Adv.<br />

Drug Deliv. Rev., 56(11), 1649–1659, 2004.<br />

25. Labhasetwar, V., Nanotechnology for drug and gene therapy: the importance of understanding molecular<br />

mechanisms of delivery, Curr. Opin. Biotechnol., 16(6), 674–680, 2005.<br />

26. Sahoo, S. K. and Labhasetwar, V., Enhanced antiproliferative activity of transferrin-conjugated<br />

paclitaxel-loaded nanoparticles is mediated via sustained intracellular drug retention, Mol. Pharm.,<br />

2(5), 373–383, 2005.<br />

27. Kommareddy, S., Tiwari, S. B., and Amiji, M. M., Long-circulating polymeric nanovectors for tumorselective<br />

gene delivery, Technol. Cancer Res. Treat., 4(6), 615–625, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


10<br />

Nanotechnology for Cancer Therapy<br />

28. Kommareddy, S. and Amiji, M., Preparation and evaluation of thiol-modified gelatin nanoparticles for<br />

intracellular DNA delivery in response to glutathione, Bioconjug. Chem., 16(6), 1423–1432, 2005.<br />

29. Kaul, G. and Amiji, M., Tumor-targeted gene delivery using poly(ethylene glycol)-modified gelatin<br />

nanoparticles: in vitro and in vivo studies, Pharm. Res., 22(6), 951–961, 2005.<br />

30. Farokhzad, O. C., Jon, S., Khademhosseini, A., Tran, T. N., Lavan, D. A., and Langer, R., Nanoparticle–aptamer<br />

bioconjugates: a new approach for targeting prostate cancer cells, Cancer Res., 64(21),<br />

7668–7672, 2004.<br />

31. Farokhzad, O. C., Khademhosseini, A., Jon, S., Hermmann, A., Cheng, J., Chin, C., Kiselyuk, A.,<br />

Teply, B., Eng, G., and Langer, R., Microfluidic system for studying the interaction of nanoparticles<br />

and microparticles with cells, Anal. Chem., 77(17), 5453–5459, 2005.<br />

32. Torchilin, V. P., Lipid-core micelles for targeted drug delivery, Curr. Drug Deliv., 2(4), 319–327,<br />

2005.<br />

33. Zeng, F., Liu, J., and Allen, C., Synthesis and characterization of biodegradable poly(ethylene glycol)block-poly(5-benzyloxy-trimethylene<br />

carbonate) copolymers for drug delivery, Biomacromolecules,<br />

5(5), 1810–1817, 2004.<br />

34. Ro<strong>by</strong>, A., Erdogan, S., and Torchilin, V. P., Solubilization of poorly soluble PDT agent, mesotetraphenylporphin,<br />

in plain or immunotargeted PEG-PE micelles results in dramatically improved cancer<br />

cell killing in vitro, Eur. J. Pharm. Biopharm., 62(3), 235–240, <strong>2006</strong>.<br />

35. Aliabadi, H. M. and Lavasanifar, A., Polymeric micelles for drug delivery, Expert Opin. Drug Deliv.,<br />

3(1), 139–162, <strong>2006</strong>.<br />

36. Rapoport, N., Combined cancer therapy <strong>by</strong> micellar-encapsulated drug and ultrasound, Int. J. Pharm.,<br />

277(1,2), 155–162, 2004.<br />

37. Gao, Z. G., Lee, D. H., Kim, D. I., and Bae, Y. H., Doxorubicin loaded pH-sensitive micelletargeting<br />

acidic extracellular pH of human ovarian A2780 tumor in mice, J. Drug Target., 13(7), 391–397,<br />

2005.<br />

38. Jeong, J. H., Kim, S. H., Kim, S. W., and Park, T. G., In vivo tumor targeting of ODN-PEG-folic<br />

acid/PEI polyelectrolyte complex micelles, J. Biomater. Sci. Polym. Ed., 16(11), 1409–1419, 2005.<br />

39. Ambade, A. V., Savariar, E. N., and Thayumanavan, S., Dendrimeric micelles for controlled drug<br />

release and targeted delivery, Mol. Pharm., 2(4), 264–272, 2005.<br />

40. Lee, J. H., Lim, Y. B., Choi, J. S., Lee, Y., Kim, T. I., Kim, H. J., Yoon, J. K., Kim, K., and Park, J. S.,<br />

Polyplexes assembled with internally quaternized PAMAM-OH dendrimer and plasmid DNA have a<br />

neutral surface and gene delivery potency, Bioconjug. Chem., 14(6), 1214–1221, 2003.<br />

41. Bhadra, D., Bhadra, S., Jain, P., and Jain, N. K., Pegnology: a review of PEG-ylated systems,<br />

Pharmazie, 57(1), 5–29, 2002.<br />

42. Asthana, A., Chauhan, A. S., Diwan, P. V., and Jain, N. K., Poly(amidoamine) (PAMAM) dendritic<br />

nanostructures for controlled site-specific delivery of acidic anti-inflammatory active ingredient,<br />

AAPS Pharm. Sci. Tech., 6(3), E536–E542, 2005.<br />

43. Khan, M. K., Nigavekar, S.S, Mine, L. D., Kariapper, M. S., Nair, B. M., Lesniak, W. G., and Balogh,<br />

L. P., In vivo biodistribution of dendrimers and dendrimer nanocomposites: Implications for cancer<br />

imaging and therapy, Technol. Cancer Res. Treat., 4(6), 603–613, 2005.<br />

44. Lesniak, W., Bielinska, A. U., Sun, K., Janczak, K. W., Shi, X., Baker, J. R., and Balogh, L. P.,<br />

Silver/dendrimer nanocomposites as biomarkers: Fabrication, characterization, in vitro toxicity, and<br />

intracellular detection, Nano Lett., 5(11), 2123–2130, 2005.<br />

45. D’Souza, G. G., Boddapati, S. V., and Weissig, V., Mitochondrial leader sequence: Plasmid DNA<br />

conjugates delivered into mammalian cells <strong>by</strong> DQAsomes co-localized with mitochondria, Mitochondrion,<br />

5(5), 352–358, 2005.<br />

46. D’Souza, G. G., Rammohan, R., Cheng, S. M., Torchilin, V. P., and Weissig, V., DQAsomemediated<br />

delivery of plasmid DNA toward mitochondria in living cells, J. Controlled Release, 92(1,2),<br />

189–197, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


2<br />

CONTENTS<br />

Passive Targeting of Solid<br />

Tumors: Pathophysiological<br />

Principles and Physicochemical<br />

Aspects of Delivery Systems<br />

S. M. Moghimi<br />

2.1 Introduction ...........................................................................................................................11<br />

2.2 Barriers to Extravasation.......................................................................................................12<br />

2.3 Selected Delivery Systems....................................................................................................12<br />

2.3.1 Liposomes..................................................................................................................12<br />

2.3.2 Polymeric Nanoparticles ...........................................................................................14<br />

2.3.3 Nanotechnology-Derived Nanoparticles ...................................................................15<br />

2.3.4 Macromolecular and Related Delivery .....................................................................16<br />

2.4 Conclusions ...........................................................................................................................16<br />

References ......................................................................................................................................17<br />

2.1 INTRODUCTION<br />

A solid tumor comprises two major cellular components: the tumor parenchyma and the stroma; the<br />

latter incorporating the vasculature and other supporting cells. As the tumor grows, in order to meet<br />

the metabolic requirements of an expanding population of tumor cells, the pre-existing blood<br />

vessels become subject to intense angiogenic pressure. Several factors produced <strong>by</strong> tumor cells<br />

and infiltrating immune-competent effector cells in the tumor parenchyma are believed to signal the<br />

development of new capillaries from the pre-existing vessels <strong>by</strong> capillary sprouting and/or dysregulated<br />

intussusceptive microvascular growth. 1 Further, in many solid tumors, endothelial cells<br />

destined to create new vessels are recruited not only from near<strong>by</strong> vessels, but also to a significant<br />

extent from precursor cells within the bone marrow (so-called endothelial progenitor cells),<br />

a process referred to as “vasculogenesis.” 2<br />

Scanning electron microscopy of microvascular corrosion casts has allowed visualization of the<br />

geometry of blood vessel architecture in solid tumors. From these studies, it has become apparent<br />

that tumor blood vessels are highly irregular and show gross architectural changes that differ from<br />

those in normal organs and from newly formed blood vessels, such as those found in wound healing<br />

and in other angiogenic sites. 1,3 For instance, the thickness of a tumor blood vessel wall is poorly<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

11


12<br />

correlated to its diameter. Therefore, despite the large size of some tumor vessels, the tumor blood<br />

flows is chaotic, with high flow rates in some segments and stagnation in others. 1 Also, the blood<br />

flow may temporarily change direction within individual tumor vessels.<br />

Further, the structure and organization of the endothelial cells, pericytes, and vascular basement<br />

membrane of tumor vessels are all abnormal. 1,3–6 One consistent abnormality of tumor blood vessels<br />

is their high permeability to macromolecules, arising from irregularly shaped and loosely interconnected<br />

endothelial cells (where the size of fenestrae often ranges from 200 to 2000 nm) and their<br />

less frequent and intimate association with pericytes and the vascular basement membrane. 1,3–5<br />

Marked variability has been noted in endothelial permeability among different tumors, different<br />

vessels within the same tumor, and during tumor growth, regression, and relapse. The extent of<br />

tumor blood vessel permeability is also controlled <strong>by</strong> the host microenvironment, and increases with<br />

the histological grade and malignant potential of tumors. 7,8<br />

Given the potency and toxicity of modern pharmacological agents, tissue selectivity is a major<br />

issue. In the delivery of chemotherapeutic agents to solid tumors this is particularly critical,<br />

since the therapeutic window for these agents is often small and the dose–response curve steep.<br />

Therefore, the idea of exploiting the well-documented vascular abnormalities of tumors, restricting<br />

penetration into normal tissue interstitium while allowing freer access to that of the tumor, becomes<br />

particularly attractive.<br />

2.2 BARRIERS TO EXTRAVASATION<br />

As a consequence of temporal and spatial heterogeneity in tumor blood flow, solid tumors usually<br />

contain well-perfused, rapidly growing regions, and poorly perfused, often necrotic areas. 1,3 As in<br />

normal tissues, diffusive and convective forces govern the movement of molecules into the interstitium<br />

of tumors. However, diffusion is believed to play a minor role in the movement of solutes<br />

across the endothelial barrier in comparison with bulk fluid flow. Examination of pressure gradients<br />

in experimental tumors has suggested that the movement of macromolecules and particulate<br />

materials out of the tumor blood vessels and into the extra-vascular compartment is remarkably<br />

limited. This has been attributed to a higher-than-expected interstitial pressure, in part due to a lack<br />

of functional lymphatic drainage, coupled with lower intra-vascular pressure. 3 In addition, interstitial<br />

pressure tends to be higher at the center of solid tumors, diminishing towards the periphery,<br />

creating a mass flow movement of fluid away from the central region of the tumor. 3 For example,<br />

the measured interstitial fluid pressure in invasive breast ductal carcinoma was 29G3 mm Hg,<br />

compared with 3.0G0.8 mm Hg in patients with benign tumors and K3.0G0.1 mm Hg in patients<br />

with normal breast parenchyma. 9 Nevertheless, the lower interstitial pressure in the periphery still<br />

permits adequate extravasation of fluid and macromolecules.<br />

These pathophysiological characteristics have serious implications for the systemic delivery of<br />

not only low-molecular-weight and macromolecular agents, but also particulate delivery vehicles.<br />

Simply enhancing the plasma half-life of these agents (e.g., long-circulating carriers) will not<br />

necessarily lead to an increase in therapeutic effect. 10 Furthermore, distribution, organization,<br />

and relative levels of collagen, decorin, and hyaluronan also impede the diffusion of extravasated<br />

macromolecules and particulate systems in tumors. 11 Thus, diffusion of macromolecules and<br />

particles will vary with tumor types, anatomical locations, and possibly <strong>by</strong> factors that influence<br />

extracellular matrix composition and/or structure.<br />

2.3 SELECTED DELIVERY SYSTEMS<br />

2.3.1 LIPOSOMES<br />

Nanotechnology for Cancer Therapy<br />

Liposomes are perhaps the best studied vehicles in cancer drug delivery, capable of either<br />

increasing the drug concentration in solid tumors and/or limiting drug exposure to critical target<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Passive Targeting of Solid Tumors 13<br />

sites such as bone marrow and myocardium. 10,12 For example, Myocete is a liposomal formulation<br />

of doxorubicin (an inhibitor of topoisomerase II) approximately 190 nm in size that was approved<br />

<strong>by</strong> the European Agency for the Evaluation of Medicinal Products (EMEA) in 2000 for the treatment<br />

of metastatic breast cancer. This formulation provides a limited degree of prolonged<br />

circulation when compared with doxorubicin in the free form. Myocete releases more than half<br />

of its associated doxorubicin within a few hours of administration and 90% within 24 h. Similar to<br />

intravenously injected nanoparticulate systems, liposomes are rapidly intercepted <strong>by</strong> macrophages<br />

of the reticuloendothelial system. 10 Hepatic deposition of Myocete could lead to gradual release of<br />

the cytotoxic agent back to the systemic circulation (a macrophage depot system), as well as<br />

induction of Kupffer cell apoptosis. 10 Following apoptosis, restoration of Kupffer cells may take<br />

up to two weeks. 13 A potentially harmful effect is the occurrence of bacteriemia during the period of<br />

Kupffer cell deficiency. Although Myocete administration decreases the frequency of cardiotoxicity<br />

and neutropenia compared with free drug, 14 there is still controversy as to whether liposomal<br />

encapsulation exhibits equivalent efficacy to doxorubicin. 15<br />

Macrophage deposition of intravenously administered liposomes can be markedly minimized<br />

either <strong>by</strong> bilayer or surface modification. 10,16 Regulatory approved examples include DaunoXome w ,<br />

a daunorubicin-encapsulated liposome 45 nm in size with a rigid bilayer for HIV-related Kaposi’s<br />

sarcoma, and Doxil w /Caelyx w , a poly(ethylene glycol)-grafted rigid vesicle of 100-nm diameter<br />

with encapsulated doxorubicin for HIV-related Kaposi’s sarcoma and refractory ovarian carcinoma.<br />

As a result of their small size, rigid bilayer, and hydrophilic surface display (as in the case of<br />

Doxil w ), these formulations exhibit poor surface opsonization, a process that limits vesicle recognition<br />

<strong>by</strong> macrophages in contact with the blood and consequently prolongs their residency time<br />

within the vasculature. 10,16 For instance, Doxil w has a biphasic circulation half-life of 84 min and<br />

46 h in humans. In addition, Doxil w also has a high drug loading capacity; here doxorubicin is loaded<br />

actively <strong>by</strong> an ammonium sulfate gradient (as doxorubicin sulfate) yielding liposomes with a high<br />

content of doxorubicin aggregates, which remain highly stable within the vasculature with minimum<br />

drug loss. 17 Therefore, it is not surprising to see that such liposomal formulations exhibit favorable<br />

pharmacokinetics when compared with the free drug. For example, the area under the curve after<br />

a dose of 50 mg/m 2 doxorubicin encapsulated in long-circulating liposomes is approximately<br />

300-fold greater than that of free doxorubicin. 17 In addition, clearance and volume of distribution<br />

are reduced <strong>by</strong> at least 250- and 60-fold, respectively. 17 However, as a result of their prolonged<br />

circulation times, alternative toxic reactions have been reported with such vehicles. The most<br />

notable dose-limiting toxicity associated with continuous infusion of Doxil w is palmar–plantar<br />

erythrodysesthesis. 17<br />

Following extravasation into solid tumors, long-circulating liposomes often distribute heterogeneously<br />

in perivascular clusters that do not move significantly and poorly interact with cancer cells. 18<br />

Therefore, the efflux of drug must follow the process of liposome extravasation at a rate that maintains<br />

free drug levels in the therapeutic range. The rate of drug release from liposomes not only depends<br />

on the composition of the interstitial fluid surrounding tumors but also on the drug type and encapsulation<br />

procedures. The importance of the latter is highlighted <strong>by</strong> the observation that extravasated<br />

long-circulating cisplatin-containing liposomes (where cisplatin is loaded passively) lack anti-tumor<br />

activity, whereas cisplatin in free form is capable of inserting cytotoxicity. 19 This is in contrast to the<br />

effective anti-tumor property of the same liposomal lipid composition containing entrapped doxorubicin.<br />

It is believed that nonspecific chemical disruption or collapse of the liposomal pH gradient, that<br />

is used to load liposomes actively with doxorubicin, may trigger doxorubicin release. 17<br />

Long-circulating liposomes have the capability to deliver between 3 and 10 times more drug to<br />

solid tumors compared with the administered drug in its free form. If the entrapped drugs are<br />

released from extravasated liposomes, it is very likely that these vesicles inherently overcome<br />

a certain degree of multidrug resistance <strong>by</strong> the tumor cells. Thus, tumor regression is to be expected<br />

with tumors exhibiting a low resistance factor. With tumors exhibiting higher resistance<br />

levels, due to over-expression of energy-dependent efflux pumps such as P-glycoprotein and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


14<br />

multidrug-resistance-related protein, alternative approaches are necessary. One effective strategy is<br />

to use long-circulating temperature-sensitive liposomes in conjugation with hyperthermia, but this<br />

approach has limited applicability for visceral and widespread malignancies. 20 Others have elaborated<br />

on biochemical triggers such as cleaveable poly(ethylene glycol)-phospholipid conjugates to<br />

generate fusion competent vesicles 21 and enzyme-mediated liposome destabilization and pore<br />

formation. 22–24 Examples of the latter include long-circulating liposomes with attached proteasesensitive<br />

haemolysin 22 and pro-drug ether liposomes (e.g., vesicles containing phospholipids with<br />

a nonhydrolyzable ether bond in the 1-position), which are susceptible to degradation <strong>by</strong> secretory<br />

phospholipases. 23,24 For instance, the level of secretory phospholipases (such as the secretory<br />

phospholipase A2) is dramatically elevated within the interstitium of various tumors.<br />

Secretory phospholiapse A2 not only acts as a trigger resulting in the release of encapsulated<br />

cytotoxic drugs from pro-drug ether liposomes, but also generates highly cytotoxic lysolipids<br />

that destabilizes plasma membrane of tumor cells, there<strong>by</strong> enhancing their permeability to cytotoxic<br />

drugs. 24<br />

There are several approaches that exploit active targeting of long-circulating liposomes to<br />

tumor cells, where receptor-mediated internalization is strongly believed to <strong>by</strong>pass tumor cell<br />

multidrug-efflux pumps. 10,16,17,25 These strategies utilize tumor-specific monoclonal antibodies<br />

or their internalizing epitopes, or ligands, such as folic acid, which are attached to the distal end<br />

of the poly(ethylene glycol) chains expressed on the surface of long-circulating liposomes. Nevertheless,<br />

with such approaches the delivery part is still passive and relies on liposome extravasation.<br />

2.3.2 POLYMERIC NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

Abraxanee is the only example of a regulatory approved (FDA, USA) nanoparticle formulation for<br />

intravenous drug delivery in cancer patients. It is paclitaxel bound to albumin nanoparticles, with a<br />

mean diameter of 130 nm, for use in individuals with metastatic breast cancer who have failed<br />

combination chemotherapy or relapse within 6 months of adjuvant chemotherapy. This formulation<br />

overcomes poor solubility of paclitaxel in the blood and allows patients to receive 50%<br />

more paclitaxel per dose over a 30-min period. 26,27 Unlike Cremophor w EL/ethanol or Tween w<br />

80-solubilized taxanes, acute hypersensitivity reactions, which are secondary to complement<br />

activation, have yet to be reported following Abraxanee infusion. Albumin nanoparticles seem<br />

to interact with gp60 receptors present on tumor blood vessels that transport the nanoparticles into<br />

tumor interstitial spaces <strong>by</strong> transcytosis, a process that may partly contribute to the effectiveness of<br />

Abraxanee. However, hepatic deposition (Kupffer cell capture) and processing of a significant<br />

fraction of albumin nanoparticles are most likely to occur. Indeed, after a 30 min infusion<br />

of 260 mg/m 2 doses of Abraxanee, faecal excretion accounted for approximately 20% of the<br />

administered dose (ABRAXIS Oncology, A division of American Pharmaceutical Partners, Inc.,<br />

Schaumburg, IL 60173, USA; 2005), thus supporting a role for hepatic handling and biliary<br />

excretion of albumin nanoparticles (or its components).<br />

Nanoparticles assembled from synthetic polymers have also received much attention in cancer<br />

drug delivery. 28 One interesting example is doxorubicin-loaded poly(alkyl cyanoacrylate) (PACA)<br />

nanoparticles. In vitro studies have indicated that PACA nanoparticles can overcome drug resistance<br />

in tumor cells expressing multidrug-resistance-1-type efflux pumps. 29 The mechanism of<br />

action is related to adherence of PACA nanoparticles to tumor cell plasma membrane, which<br />

initiates particle degradation and provides a concentration gradient for doxorubicin, and diffusion<br />

of doxorubicin across the plasma membrane following formation of an ion pair between the<br />

positively charged doxorubicin and the negatively charged cyanoacrylic acid (a nanoparticle<br />

degradation product). 29 These observations clearly indicate that drug release and nanoparticle<br />

degradation must occur simultaneously, yielding an appropriate size complex with correct physicochemical<br />

properties for diffusion across the plasma membrane. Further developments with<br />

PACA nanoparticles include preparations that contain doxorubicin within the particle core and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Passive Targeting of Solid Tumors 15<br />

cyclosphorin, an inhibitor of the P-glycoprotein, at the surface. 30 Similar to liposomes, long-circulating<br />

versions of PACA nanoparticles have also been engineered for passive as well as active<br />

targeting to solid tumors. 31<br />

2.3.3 NANOTECHNOLOGY-DERIVED NANOPARTICLES<br />

Nanotechnology is a cross-disciplinary field, which involves the ability to design and exploit the<br />

unique properties that emerge from man-made materials ranging in size from 1 to greater than<br />

100 nm. 16 Indeed, the physical and chemical properties of materials—such as porosity, electrical<br />

conductivity, light emission, and magnetism—can significantly improve or radically change<br />

as their size is scaled down to small clusters of atoms. These advances are beginning to have<br />

a paradigm-shifting impact not least in experimental (e.g., thermal tumor killing) and diagnostic<br />

oncology. 16,32 Examples include superparamagnetic iron oxide nanocrystals, quantum dots (QDs),<br />

inorganic nanoparticles, and composite nanoshells. The surfaces of these entities are amenable to<br />

modification with synthetic polymers (to afford long-circulating properties) and/or to targeting<br />

ligands. However, a key problem with these technologies is toxicity and is discussed elsewhere. 16<br />

Iron oxide nanocrystals are formed from an inner core of hexagonally shaped iron oxide<br />

particles of approximately 5 nm, which express correlated electron behavior; at a high enough<br />

temperature, they are superparamagnetic. 33,34 In addition, dextran or synthetic polymers such as<br />

poly(ethyleneglycol) surround the crystal core. Indeed, it is the combination of the small size and<br />

surface characteristics that allow iron oxide nanocrystals, once injected into the blood stream, to<br />

<strong>by</strong>pass rapid detection <strong>by</strong> the body’s defence cells and accumulate in tumor sites <strong>by</strong> extravasation.<br />

Therefore, they are useful for patient selection, detection of tumor progression, and tracking of the<br />

effectiveness of anti-tumor treatment regimens <strong>by</strong> magnetic resonance imaging (MRI). These<br />

approaches can be extended for site-specific imaging of tumor vasculature with targeting<br />

ligands. In addition, iron oxide nanocrystals can slowly extravasate from the vasculature into the<br />

interstitial spaces, from which they are transported to lymph nodes <strong>by</strong> way of lymphatic vessels. 34<br />

Within lymph nodes they are captured <strong>by</strong> local macrophages, and their intracellular accumulation<br />

shortens the spin relaxation process of near<strong>by</strong> protons detectable <strong>by</strong> MRI. On magnetic resonance<br />

images, those node regions accumulating iron oxide appear dark relative to surrounding tissues.<br />

Indeed, iron oxide nanocrystals can distinguish between normal and tumor-bearing nodes and<br />

reactive and metastatic nodes. 34<br />

QDs are made of semiconductors like silicon and gallium arsenide. 35,36 In these particles there<br />

are discrete electronic energy levels (valance band and conduction band), but the spacing of the<br />

electronic energy levels (band gap) can be precisely controlled through variation in size. When<br />

a photon, with higher energy than the energy of the band gap, hits a QD, an electron is promoted<br />

from valance band into the conduction band, leaving a hole behind. Electrons emit their excess<br />

energy as light when they recombine with holes. Since optical response is due to the excitation<br />

of single electron-hole pairs, the size and shape of QDs can be tailored to fluoresce specific colors.<br />

The ability of QDs to tune broad wavelength together with their photostability is of paramount<br />

importance in biological labeling. 35,36 Indeed, QDs stay lit much longer than conventional dyes<br />

used for imaging and tagging purposes and therefore have the potential to improve the resolution of<br />

tumor cells to the single cell level <strong>by</strong> optical imaging as well as determining heterogeneity among<br />

cancer cells in a solid tumor. 37–39 Unlike QDs, where optical response is due to the excitation of<br />

single electron hole pairs, in metallic nanoparticles (e.g., gold) incident light can couple to the<br />

plasmon excitation of the metal. This involves the light-induced motion of all valence electrons. 36<br />

Therefore, the type of plasmon that exists on a surface of a metallic nanoparticle is directly related<br />

to its shape and curvature; so it is possible to make a wide range of light scatterers that can be<br />

detected at different wavelengths. Composite nanoshells consist of a spherical dielectric core (e.g.,<br />

silica) surrounded <strong>by</strong> a thin metal shell (e.g., gold). Again, <strong>by</strong> controlling the relative thickness of<br />

the core and shell layers of the composite nanoparticle, the plasmon resonance and the resultant<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


16<br />

optical absorption properties can be tuned from near-UV to the mid-infrared. Of particular interest<br />

is the ability of near-infrared light (700–1000 nm) to penetrate through tissue at depths of a few cm<br />

with minimal heat generation and tissue damage. Thus, a recent study demonstrated rapid irreversible<br />

photothermal ablation of tumor tissue in vivo following administration of near-infraredabsorbing<br />

silica–gold nanoshells in combination with an extracorporeal low-power diode laser. 40<br />

2.3.4 MACROMOLECULAR AND RELATED DELIVERY<br />

Polymer-based drug delivery systems also favorably alter the pharmacokinetics and biodistribution<br />

of conjugated drugs and accumulate in tumor interstitium following extravasation. 41 Examples<br />

include SMANCS (a conjugate of the polymer styrene-co-maleic acid/anhydride and neocarzinostatin<br />

for treatment of hepatocellular carcinoma), conjugates of various cytotoxic agents (e.g.,<br />

paclitaxel, doxorubicin, platinate, and campthothecin) with polyglutamate and nonbiodegradable<br />

hydroxypropyl methacrylamide.<br />

Other related polymer-based systems in cancer drug delivery include micelles 42,43 and dendrimers.<br />

44 For example, Pluronics w (copolymers of ethylene oxide and propylene oxide) are capable<br />

of forming micelles, and some members of Pluronic copolymers can overcome multidrug resistance.<br />

42 However, it is becoming clear that Pluronic copolymers can induce complement activation,<br />

even at concentrations below their critical micelle concentration, which may increase the risk<br />

of pseudoallergy in sensitive patients. 45<br />

Dendrimers are highly branched macromolecules with controlled near monodisperse threedimensional<br />

architecture emanating from a central core. 44 Polymer growth starts from a central<br />

core molecule and growth occurs in an outward direction <strong>by</strong> a series of polymerization reactions.<br />

Hence, precise control over size can be achieved <strong>by</strong> the extent polymerization, starting from a few<br />

nanometers. Cavities in the core structure and folding of the branches create cages and channels.<br />

The surface groups of dendrimers are amenable to modification and can be tailored for specific<br />

applications. Therapeutic and diagnostic agents are usually attached to surface groups on dendrimers<br />

<strong>by</strong> chemical modification. For example, a recent study has used tagged-dendrimers for in vivo<br />

evaluation of tumor-associated matrix metalloproteinase-7 (matrilysin) activity. 46<br />

Other macromolecular systems for cancer targeting and treatment include various forms of<br />

monoclonal and bispecific monoclonal antibodies against tumor-associated antigens. 47 These can<br />

further be coupled to drugs, toxins, enzymes (as in antibody-directed enzyme pro-drug therapy),<br />

cytokines, radionuclides, etc.<br />

2.4 CONCLUSIONS<br />

Nanotechnology for Cancer Therapy<br />

The chaotic blood flow in tumor vasculature and the heterogeneous vascular permeability of tumor<br />

blood vessels are among the key barriers controlling passive delivery of macromolecular and<br />

particulate delivery systems into the interstitium of solid tumors. Already compromised <strong>by</strong><br />

abnormal hydrostatic pressure gradients, compressive mechanical forces generated <strong>by</strong> tumor cell<br />

proliferation cause intratumoral vessels to compress and collapse, thus creating further barriers for<br />

passive targeting. Interestingly, tumor-specific cytotoxic therapy, reducing tumor cell number, may<br />

result in more efficient delivery <strong>by</strong> decompressing these same vessels, 48 but this enhanced perfusion<br />

could provide a route for metastasis. Pathohysiologoical barriers, however, are not fully developed<br />

in micrometastases, and also pose a lesser problem in the diagnostic oncology as well as in drug<br />

delivery to well-perfused and low-pressure regions in larger tumors. Some of the problems may<br />

possibly be overcome <strong>by</strong> design of long-circulating multifunctional carriers (carriers that contain<br />

appropriate combinations of cytotoxic agents, diagnostic, and barrier-avoiding components) with<br />

biochemical triggering mechanisms. 16 The vascular barrier of the solid tumor is also its Achilles’<br />

heel; the nutritionally demanding tumor cells are entirely dependent upon a functional vasculature.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Passive Targeting of Solid Tumors 17<br />

For this reason, interest has also been focused on the concept of tumor vasculature as a target rather<br />

than a barrier and is reviewed in this compendium and elsewhere. 49–51<br />

REFERENCES<br />

1. Munn, L. L., Aberrant vascular architecture in tumors and its importance in drug-based therapies,<br />

Drug Discov. Today, 8, 396, 2003.<br />

2. Lyden, D. et al., Impaired recruitment of bone-marrow-derived endothelial and haematopoietic<br />

precursor cells blocks, tumor angiogenesis, and growth, Nat. Med., 7, 1194, 2001.<br />

3. Jain, R. K., Delivery of molecular medicine to solid tumors: Lessons from in vivo imaging of gene<br />

expression and function, J. Control. Release, 74, 7, 2001.<br />

4. Dvorak, H. F. et al., Identification and characterization of the blood vessels of solid tumors that are<br />

leaky to circulating macromolecules, Am. J. Pathol., 133, 95, 1988.<br />

5. Hashizume, H. et al., Openings between defective endothelial cells explain tumor vessel leakiness,<br />

Am. J. Pathol., 156, 1363, 2000.<br />

6. Morikawa, S. et al., Abnormalities in pericytes on blood vessels and endothelial sprouts in tumors,<br />

Am. J. Pathol., 160, 985, 2002.<br />

7. Fukumura, D. et al., Effect of host microenvironment on the microcirculation of human colon adenocarcinoma,<br />

Am. J. Pathol., 151, 679, 1997.<br />

8. Daldrup, H. et al., Correlation of dynamic contrast-enhanced MR imaging with histologic tumor<br />

grade: Comparison of macromolecular and small-molecular contrast media, Am. J. Roentgenol.,<br />

171, 941, 1998.<br />

9. Nathanson, S. D. and Nelson, L., Interstitial fluid pressure in breast cancer, benign breast conditions,<br />

and breast parenchyma, Ann. Surg. Oncol., 1, 333, 1994.<br />

10. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

Theory to practice, Pharmacol. Rev., 53, 283, 2001.<br />

11. Pluen, A. et al., Role of tumor–host interactions in interstitial diffusion of macromolecules: Cranial vs.<br />

subcutaneous tumors, Proc. Natl Acad. Sci. USA, 98, 4628, 2001.<br />

12. Torchilin, V. P., Recent advances with liposomes as pharmaceutical carriers, Nat. Rev. Drug Discov.,<br />

4, 145, 2005.<br />

13. Daemen, T. et al., Liposomal doxorubicin-induced toxicity: Depletion and impairment of phagocytic<br />

activity of liver macrophages, Int. J. Cancer, 61, 716, 1995.<br />

14. Batist, G. et al., Reduced cardiotoxicity and preserved antitumor efficacy of liposome-encapsulated<br />

doxorubicin and cyclophosphamide compared with conventional doxorubicin and cyclophosphamide<br />

in a randomized, multicenter trial of metastatic breast cancer, J. Clin. Oncol., 19, 1444, 2001.<br />

15. Williams, G., Cortazar, P., and Pazdur, R., Developing drugs to decrease the toxicity of<br />

chemotherapy, J. Clin. Oncol., 19, 3439, 2001.<br />

16. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Nanomedicine: Current status and future prospects,<br />

FASEB J., 19, 311, 2005.<br />

17. Gabizon, A., Shmeeda, H., and Barenholz, Y., Pharmacokinetics of pegylated liposomal doxorubicin:<br />

Review of animal and human studies, Clin. Pharmacokinet., 42, 419, 2003.<br />

18. Yuan, F. et al., Microvascular permeability and interstitial penetration of sterically stabilized (stealth)<br />

liposomes in a human tumor xenograft, Cancer Res., 54, 3352, 1994.<br />

19. Bandak, S. et al., Pharmacological studies of cisplatin encapsulated in long-circulating liposomes in<br />

mouse tumor models, Anti-Cancer Drugs, 10, 911, 1999.<br />

20. Huang, S. K. et al., Liposomes and hyperthermia in mice: Increased tumor uptake and therapeutic<br />

efficacy of doxorubicin in sterically stabilized liposomes, Cancer Res., 54, 2186, 1994.<br />

21. Zalipsky, S. et al., New detachable poly(ethylene glycol) conjugates: Cysteine–cleavable lipopolymers<br />

regenerating natural phospholipid, diacylphosphatidylethanolamine, Bioconjug. Chem., 10, 703,<br />

1999.<br />

22. Provoda, C. J. and Lee, K.-D., Bacterial pore-forming hemolysins and their use in the cytosolic<br />

delivery of macromolecules, Adv. Drug Deliv. Rev., 41, 209, 2000.<br />

23. Andresen, T. L., Jensen, S. S., and Jørgensen, K., Advanced strategies in liposomal cancer therapy:<br />

Problems and prospects of active and tumor specific drug release, Prog. Lipid Res., 44, 68, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


18<br />

Nanotechnology for Cancer Therapy<br />

24. Andresen, T. L. et al., Triggered activation and release of liposomal prodrugs and drugs in cancer<br />

tissue <strong>by</strong> secretory phospholipase A2, Curr. Drug Deliv., 2, 353, 2005.<br />

25. Allen, T. M., Ligand-targeted therapeutics in anti-cancer therapy, Nat. Rev. Cancer, 2, 750, 2002.<br />

26. Bartels, C. L. and Wilson, A. F., How does a novel formulation of paclitaxel affect drug delivery in<br />

metastatic breast cancer?, US Pharm., 29, HS18, 2004.<br />

27. Garber, K., Improved paclitaxel formulation hints at new chemotherapy approach, J. Natl Cancer<br />

Inst., 96, 90, 2004.<br />

28. Brigger, I., Dubernet, C., and Couvreur, P., Nanoparticles in cancer therapy and diagnosis, Adv. Drug<br />

Deliv. Rev., 54, 631, 2002.<br />

29. Colin de Verdiere, A. et al., Reversion of multidrug resistance with polyalkylcyanoacrylate nanoparticles:<br />

Towards a mechanism of action, Br. J. Cancer, 76, 198, 1997.<br />

30. Soma, C. E. et al., Reversion of multidrug resistance <strong>by</strong> co-encapsulation of doxorubicin and cyclosporin<br />

A in polyalkylcyanoacrylate nanoparticles, Biomaterials, 21, 1, 2000.<br />

31. Stella, B. et al., Design of folic acid-conjugated nanoparticles for drug targeting, J. Pharm. Sci., 89,<br />

1452, 2000.<br />

32. Ferrari, M., Cancer nanotechnology: Opportunities and challenges, Nat. Rev. Cancer, 5, 161, 2005.<br />

33. Perez, J. M., Josephson, L., and Weissleder, R., Use of magnetic nanoparticles as nanosensors to probe<br />

for molecular interactions, Chem. Bio. Chem., 5, 261, 2004.<br />

34. Moghimi, S. M. and Bonnemain, B., Subcutaneous and intravenous delivery of diagnostic agents to<br />

the lymphatic system: Applications in lymphoscintigraphy and indirect lymphography, Adv. Drug<br />

Deliv. Rev., 37, 295, 1999.<br />

35. Medintz, I. L. et al., Quantum dot bioconjugates for imaging, labelling, and sensing, Nat. Mater.,<br />

4, 435, 2005.<br />

36. Alivisatos, P., The use of nanocrystals in biological detection, Nat. Biotechnol., 22, 47, 2003.<br />

37. Stroh, M. et al., Quantum dots spectrally distinguish multiple species within the tumor milieu in vivo,<br />

Nat. Med., 11, 678, 2005.<br />

38. Wu, X. et al., Immunofluorescent labeling of cancer marker Her2 and other cellular targets with<br />

semiconductor quantum dots, Nat. Biotechnol., 21, 41, 2003.<br />

39. Gao, X. et al., In vivo cancer targeting and imaging with semiconductor quantum dots, Nat.<br />

Biotechnol., 22, 969, 2004.<br />

40. Hirsch, L. R. et al., Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic<br />

resonance guidance, Proc. Natl Acad. Sci. USA, 100, 13549, 2003.<br />

41. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2, 347, 2003.<br />

42. Oh, K. T., Bronich, T. K., and Kabanov, A. V., Micellar formulations for drug delivery based on<br />

mixtures of hydrophobic and hydrophilic Pluronic w block copolymers, J. Control. Release, 94, 411,<br />

2004.<br />

43. Adams, M. L., Lavasanifar, A., and Kwon, G. S., Amphiphilic block copolymers for drug delivery,<br />

J. Pharm. Sci., 92, 1343, 2003.<br />

44. Haag, R., Supramolecular drug-delivery systems based on polymeric core-shell architectures, Angew.<br />

Chem. Int. Ed., 43, 278, 2004.<br />

45. Moghimi, S. M. et al., Causative factors behind poloxamer 188 (Pluronic F68 Flocor)-induced complement<br />

activation in human sera: a protective role against poloxamer-mediated complement activation<br />

<strong>by</strong> elevated serum lipoprotein levels, Biochim. Biophys. Acta-Mol. Basis Dis., 103, 1689, 2004.<br />

46. McIntyre, J. O. et al., Development of a novel fluorogenic proteolytic beacon for in vivo detection<br />

and imaging of tumor-associated matrix metalloproteinase-7 activity, Biochem. J., 377, 617, 2004.<br />

47. Kosterink, J. G. W. et al., Strategies for specific drug targeting to tumor cells. Vol. 12 of Drug Targeting,<br />

Organ-Specific Strategies, Molema, G., and Meijer, D. K. F., Eds., Wiley-VCH, Weinheim pp.199–232,<br />

2001, chapter 8.<br />

48. Padera, T. P. et al., Cancer cells compress intratumor vessels, Nature, 427, 695, 2004.<br />

49. Pasqualini, R., Arap, W., and McDonald, D. M., Probing the structural and molecular diversity of<br />

tumor vasculature, Trend Mol. Med., 8, 563, 2002.<br />

50. Murray, J. C. and Moghimi, S. M., Endothelial cells as therapeutic targets in cancer: New biology<br />

and novel delivery systems, Crit. Rev. Ther. Drug Carrier Syst., 20, 139, 2003.<br />

51. Reynolds, A. R., Moghimi, S. M., and Hodivala-Dilke, K., Nanoparticle-mediated gene delivery<br />

to tumor neovasculature, Trend Mol. Med., 9, 2, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


3<br />

CONTENTS<br />

Active Targeting Strategies in<br />

Cancer with a Focus on Potential<br />

Nanotechnology Applications<br />

Randall J. Mrsny<br />

3.1 Introduction ........................................................................................................................... 19<br />

3.2 Nanoparticle Characteristics ................................................................................................. 20<br />

3.2.1 Composition and Biocompatibility ........................................................................... 22<br />

3.2.2 Derivitization .............................................................................................................23<br />

3.2.3 Detection.................................................................................................................... 23<br />

3.3 Nanoparticle Targeting.......................................................................................................... 24<br />

3.3.1 Inherent Targeting ..................................................................................................... 24<br />

3.3.2 Complicating Aspects ............................................................................................... 25<br />

3.3.3 Safety Issues .............................................................................................................. 26<br />

3.4 Targeting to Cancer Cells ..................................................................................................... 27<br />

3.4.1 Cell Surface Properties.............................................................................................. 28<br />

3.4.2 Metabolic Properties ................................................................................................. 29<br />

3.5 Targeting to Tumors.............................................................................................................. 30<br />

3.5.1 Aberrant Vasculature................................................................................................. 30<br />

3.5.2 Metabolic Environment............................................................................................. 32<br />

3.6 Fate of Nanoparticles ............................................................................................................ 32<br />

3.6.1 Site of Initial Application ......................................................................................... 33<br />

3.6.2 Distribution to Specific Organs................................................................................. 34<br />

3.6.3 Intracellular Uptake and Fate.................................................................................... 34<br />

3.7 Conclusions ........................................................................................................................... 35<br />

References ...................................................................................................................................... 37<br />

3.1 INTRODUCTION<br />

Nanotechnology has recently become a buzzword in several scientific fields, including the area<br />

of drug delivery, 1 and a variety of nanomedicine opportunities have recently been reviewed. 2 The<br />

concept of nanoparticles as drug delivery vehicles is not new as reviews on the subject were published<br />

nearly 20 years ago. 3 That nanoparticles could be selectively targeted <strong>by</strong> coating with monoclonal<br />

antibodies provided an early, ground-breaking proof-of-concept that these materials might one day<br />

be effective tools for the diagnosis and treatment of cancers. At their inception, all new technologies<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

19


20<br />

Targeting<br />

Imaging<br />

appear to have a plethora of possibilities because limitations have not yet become apparent. Subsequent<br />

studies that examine the limits of various applications then provide a menu of feasible<br />

applications. This menu can change as new developments of the technology are realized that<br />

allow one to address additional applications. This has certainly been the case for nanotechnology<br />

applications that involve active targeting to cancers for diagnostic and therapeutic purposes. 4<br />

Although nanoparticles have shown tremendous promise in facilitating the targeted delivery of<br />

therapeutics and diagnostics to cancers, the composition and size of the particles have inherent<br />

physical and chemical properties that can compromise their capability to localize to and/or treat<br />

cancers because of non-selective cell and tissue uptake. Therefore, active targeting to cancers using<br />

nanoparticles requires consideration of not only unique properties of the cancer that allow for<br />

specific targeting but also attention to issues that minimize non-selective delivery to uninvolved<br />

regions of the body. Methods to overcome functional barriers that limit uptake of materials or<br />

delivery to cancer cells must be also considered. Even with proper consideration of these issues,<br />

successful disposition and targeting to cancers is quite a challenge. Certainly, lack of consideration<br />

of these issues can dramatically increase the risk of serious negative outcomes, particularly when<br />

targeted nanoparticles contain potent cytotoxic agents.<br />

Advantages and disadvantages of specific nanoparticles as well as methods to potentially<br />

correct shortcomings of nanoparticle targeting to cancers will be discussed in this chapter.<br />

Several of the topics raised in this chapter will also be discussed in much greater depth in other<br />

chapters in this text. Specifically, two major strategies of cancer targeting related to nanotechnology<br />

opportunities will be addressed: targeting to cancer cells and targeting to tumors. Attention will be<br />

paid to similarities as well as differences for these two strategies. Targeting cancer cells and tumors<br />

for diagnostic purposes as well as therapy will also be examined. Several applications for nanoparticles<br />

have been outlined <strong>by</strong> the National Cancer Institute (http://nano.cancer.gov) regarding<br />

unmet medical needs in the areas of targeting, imaging, delivery, and reporting agents for cancer<br />

diagnosis and treatment (Figure 3.1). General cellular responses to and fates of nanoparticles that<br />

can affect these potential applications will be discussed as critical aspects of ultimate<br />

clinical success.<br />

3.2 NANOPARTICLE CHARACTERISTICS<br />

Delivery<br />

Reporting<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 3.1 Integrated potential applications of nanoparticles as visualized <strong>by</strong> the National Cancer Institute<br />

of the NIH. Nanoparticles, because of their general capacity for multiple modifications as well as their inherent<br />

properties, can have multiple functions relevant to targeting cancers for therapeutic and/or diagnostic purposes.<br />

Clinical success of nanoparticle-based diagnostics and therapeutics requires proper matching of<br />

particle characteristics. The characteristics of nanoparticles are critically dependent upon the<br />

materials used to prepare the nanoparticle. Nanoparticles can now be readily prepared from a<br />

wide range of inorganic and organic materials in a range of sizes from two to several hundred<br />

nanometers (nm) in diameter. Put in perspective, human cells are typically 10,000–20,000 nm in<br />

diameter. The plasma membrane of these cells is 6 nm in thickness. In most cases, nanoparticles<br />

can be generated to have narrow and defined size ranges. Other chapters in this text will focus on the<br />

physical and chemical characteristics of nanoparticles made from various materials as well as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 21<br />

methods for their production. Although nanoparticles can be prepared from a wide variety of<br />

materials (inorganic salts, lipids, synthetic organic polymers, polymeric forms of amino acids,<br />

nucleic acids, etc.), this chapter will primarily focus on those prepared from materials that would<br />

be considered sufficiently safe for repeated systemic administrations and/or would be perceived to<br />

have an acceptable safety profile that would warrant use in man. In general, it is desirable for<br />

nanoparticles to be either readily metabolized or sufficiently broken down to produce only nontoxic<br />

metabolites that can be safely excreted. Indeed, tremendous advances have been made in<br />

controlling the chemical nature, degradable characteristics, and dimensions of nanoparticles.<br />

Many of the initial studies examining nanoparticles as delivery tools used particles prepared<br />

from materials such as polyalkylcyanoacrylates (PAA). 5 The extreme stability of PAA is both a<br />

positive and a negative. PAA nanoparticles will not be degraded prior to reaching a tissue or cell<br />

target site; however, once they reach that site, it is unlikely that they will be efficiently metabolized.<br />

Therefore, PAA nanoparticles have been extremely useful for initial studies of nanomaterials for<br />

cancer targeting, but an inability to clear PAA nanoparticles presents uncertainties as to their<br />

ultimate toxicological fate. Concerns over repeated PAA nanoparticle administrations in man<br />

and the need for more acceptable materials were highlighted early on. 3 One of the biggest concerns<br />

regarding poorly metabolized nanoparticles is that of accumulation and the potential sequelae<br />

associated with such an outcome. In some cases where a limited number of exposures would<br />

occur, one could consider the use of materials that are not readily metabolized <strong>by</strong> the body. In<br />

the case of certain cancer applications, it might be possible to use materials that otherwise would be<br />

considered to have an unacceptable safety signal following repeat dosing or that have the potential<br />

to accumulate. Therefore, rationales exist for the potential application of nanoparticles prepared<br />

from a wide range of materials, even those that, at first glance, would be considered unacceptable.<br />

Methods of production and composition define nanoparticle characteristics; these characteristics<br />

define potential issues (and opportunities) related to biocompatibility, derivitization, and<br />

detection. Nanoparticles can be prepared from a singular subunit that is chemically coupled and<br />

organized in a defined (e.g., dendrimers) or in a more random (e.g., polylactic acid) manner.<br />

Although these materials would not have a defined core, they can be impregnated with compatible<br />

materials and/or chemically modified at their surface. Materials such as glyconanoparticles would<br />

provide one approach where a distinct core with radiating ligands could be positioned using linkers.<br />

In such a case, the solid core, used to anchor each linker used for the attachment of targeting<br />

ligands, could be used to deliver a therapeutic or diagnostic payload. Liposomes are an example of<br />

nanoshell structures that can be loaded internally as well as impregnated within the shell. Many<br />

types of nanomaterials fall into one of these three general structural architectures (Figure 3.2).<br />

(a) (b) (c)<br />

FIGURE 3.2 General schema for three types of nanoparticle structures. (a) Nanoparticles can be formed from<br />

one type of material that can be impregnated with therapeutic or lipophilic imaging reagents (open diamonds)<br />

and modified with targeting ligands (crescents) positioned <strong>by</strong> chemical coupling through linker moieties. (b)<br />

Metal (or similar) cores (circles can be modified through a linker-targeting ligand system to generate another<br />

type of nanoparticle structure. In this case, it might be possible to use elaborated linkers as an environment<br />

compatible for incorporation of therapeutic or imaging reagents. (c) Shell-type nanoparticles such as liposomes<br />

where an aqueous compartment is enclosed <strong>by</strong> a bilayer of phospholipids can also be used for the targeted<br />

delivery of hydrophilic therapeutic or imaging reagents (filled hexagons).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


22<br />

3.2.1 COMPOSITION AND BIOCOMPATIBILITY<br />

Nanotechnology for Cancer Therapy<br />

Biological organic polymers can be formed from amino acids to form peptides and proteins. Some<br />

proteins have, <strong>by</strong> themselves, been used as nanoparticle delivery systems. Indeed, the protein<br />

ferritin functions as a coated nanoparticle. This molecule is w12 nm in diameter and can carry<br />

hydrous ferrous oxide (w5–7 nm in diameter) during its role as an iron storage system for the<br />

body. 6 Alternately, proteins can be used to generate nanoparticles for carrier applications; gelatin<br />

has been used to prepare nanoparticles. 7 Albumin nanoparticles have been described. 8 A 13-MDa<br />

ribonucleoprotein, termed a vault, has also been identified as a nanomaterial that could be used to<br />

deliver therapeutic and diagnostic agents. 9 Materials prepared from nucleic acids also have the<br />

potential to act as nanoparticle carriers. In general, nanoparticles prepared from biological materials<br />

would be biocompatible as a result of obvious elimination mechanisms.<br />

Nanoparticles can also be prepared from biological materials that are found in the body but<br />

are not typically organized as nanoparticle-size polymers; complex mixtures of polysaccharides,<br />

poly-lysine, and poly(D,L-lactic and glycolic acids; PLGA) have been prepared in a variety of sizes<br />

and in formats that allow ligand coupling with targeting moieties as well as diagnostic or therapeutic<br />

agents. 10 Synthetic organic polymers such as PLGA have been used to produce resorbable<br />

sutures, providing nanoparticles that will produce a sustained release of its contents. PLGA nanoparticles<br />

have been used to deliver wild-type p53 protein to cancer cells. 11 Such polymers have<br />

been studied for the development of nanoparticle delivery vehicles. 12–14 Although such nanoparticles<br />

would be considered relatively safe because of their biocompatibility, particles prepared from<br />

these types of materials can initiate inflammatory responses at sites of accumulation or deposit. 15<br />

A number of new synthetic organic polymer materials are being examined for their capacity to<br />

generate nanoparticles useful for drug encapsulation and targeting. 16 In general, organic polymerbased<br />

nanoparticles have provided promising results in the field of cancer targeting for diagnostics<br />

and therapy with the added capability of sustained release in some cases. 17<br />

Dendrimer structures are prepared from a series of repetitive chemical steps that perpetually<br />

increase their size as additional shells are added. Varying the seed molecule can produce nanoparticle<br />

structures of spherical or more elongated shapes. Dendrimer-based nanoparticle structures<br />

are inherently different from other organic-based nanoparticles prepared as linear sequences<br />

of subunits (e.g., amino acids used to form proteins) or from subunits that can randomly branch<br />

(e.g., polysaccharides). With the flexibility of chemistries that have now been described to prepare<br />

dendrimers, these materials show tremendous promise to prepare highly defined nanoparticles that<br />

can be targeted to cancers for therapeutic and diagnostic applications. To date, polyamidoamine<br />

(PAMAM) dendrimers have been the most extensively studied family of dendrimers that can be<br />

used to deliver anti-cancer drugs. 18<br />

A variety of inorganic materials can be used to generate nanoparticles for drug delivery that<br />

might be applied to cancer and tumor targeting. For example, iron oxide (Fe2O3) nanoparticles can<br />

be used to deliver anti-cancer agents. 19 Fe2O3 nanoparticles targeted to a cancer can become hot<br />

enough in an applied oscillating magnetic field to kill cells. Calcium phosphate precipitates can also<br />

be also made into nanoparticles. Although calcium phosphate precipitates can be metabolized over<br />

time and would be considered biocompatible, these materials can act as a potent adjuvant, potentially<br />

enhancing their application to target the delivery of cancer antigens. 20 In this way, calcium<br />

phosphate nanoparticles are similar to another inorganic salt precipitate, aluminum hydroxide<br />

(alum), that is currently approved as an adjuvant for human vaccines. 21 Semiconductor nanocrystals<br />

(quantum dots) are another example of inorganic nanoparticles. Quantum dots have exceptional<br />

characteristics for in vivo imaging and diagnostic applications. 22,23 However, some materials used<br />

to generate quantum dots such as CdSe can release toxic Cd 2C ions that alter ion channel function<br />

and lead to cell death when sufficient levels are reached. 24 Therefore, some of the inorganic<br />

materials used to generate nanoparticles may have significant biocompatibility issues.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 23<br />

Lipids such as phospholipids and cholesterol can be used to generate single- or multi-lamellar<br />

spheres (liposomes) in the nanometer-size range. Liposome-based nanocapsules that can be loaded<br />

with diagnostic or therapeutic agents for the targeted delivery to cancers 25 and enhancement<br />

strategies for lipid-based nanoparticle-mediated tumor targeting have been reviewed. 26 Liposomes<br />

were some of the first nanoparticle structures extensively evaluated for cancer targeting. Early<br />

studies using liposomes highlighted issues associated with recognition and clearance <strong>by</strong> cells of the<br />

reticuloendothelial system (RES) that remove particulates from the systemic circulation. 27 Methods<br />

of masking liposomes from the RES such as modification with poly(ethylene glycol) (PEG) have<br />

been successfully used to limit RES clearance and increase circulating half-lives in serum. 28<br />

Solid lipid nanoparticles, nanostructure lipid carriers, and lipid–drug conjugate nanoparticles<br />

have also been described for the drug delivery strategies that could be applied to cancer diagnosis<br />

and/or therapy. 29 Lipid-based nanospheres can be sterically stabilized <strong>by</strong> the incorporation of<br />

artificial lipid derivatives that can be cross-linked. Subsequently, stabilized lipid-based nanospheres<br />

can be targeted to cancers using a conjugated antibody. 30 Such covalent modifications can improve<br />

the stability of lipid-based nanoparticles but can also reduce the biocompatible natures of these<br />

materials <strong>by</strong> modifying their clearance from the body. Apolipoprotein E-containing liposomes have<br />

also been prepared as a carrier for a lipophilic prodrug of daunorubicin as a means of targeting<br />

cancer cells that overexpress the receptor for low-density lipoproteins (LDL). 31<br />

3.2.2 DERIVITIZATION<br />

Because of their chemical and physical characteristics, nanoparticles exhibit inherent cellular<br />

targeting and uptake characteristics. Size and surface charge seem to be the two prominent characteristics<br />

that affect inherent nanoparticle targeting and cellular uptake. Because inherent targeting<br />

mechanisms may not provide the targeting or delivery characteristics desired, methods to modify<br />

nanoparticles with targeting agents can be critical. Although some nanoparticle materials are<br />

composed of materials with functional groups useful for chemical coupling, others are not. Such<br />

nanoparticle systems must be either modified to allow chemical coupling or doped with reagents<br />

that can be used for this modification. A number of coupling strategies have also been worked out<br />

that allow for efficient functionalization of nanomaterials through both reversible and irreversible<br />

chemistries. 14,32 These modifications allow for the coupling of antibodies, receptor ligands, and<br />

other potential targeting agents. Similar to the concerns associated with composition of the nanoparticle<br />

itself, any modification through chemical derivitization must also be considered with regard<br />

to generating materials with unacceptable toxicity or neutralization of the function of the nanoparticle<br />

or its targeting element.<br />

Nanoparticles have the advantage that they can be modified with multiple ligands to enhance<br />

their targeting selectivity and/or allow for simultaneous delivery of diagnostic and therapeutic<br />

agents. 33 It is important to appreciate the relative size of components used to construct and derivitize<br />

nanoparticles. For example, a quantum dot may be only 10 nm in diameter. Targeting that<br />

sized particle with an antibody might require the attachment of an IgG antibody that is roughly<br />

equal in size. By comparison, a fluorescent material that might be useful for localization of a<br />

targeted nanoparticle such as green fluorescent protein (GFP) is about 5 nm. Derivitization<br />

strategies for the construction of targeted nanoparticles must incorporate a consideration of<br />

potential steric conflicts for incorporation of targeting, detection, and therapeutic components.<br />

Other chapters in this text will extensively examine derivitization technologies for nanoparticles.<br />

3.2.3 DETECTION<br />

Most, if not all, nanoparticle structures currently investigated for delivery of cancer therapeutics<br />

also have the capacity to be detected or modified to contain a detectable agent that could be<br />

simultaneously used for cancer diagnosis. For example, PAMAM folate-dendrimers that contain<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


24<br />

contrast media for detection <strong>by</strong> magnetic resonance imaging (MRI) are effectively targeted to<br />

cancer cells that overexpress the high affinity folate receptor. 34 Gadolinium ion–dendrimer nanoparticlesarereadilydetected<strong>by</strong>CTimagingand<br />

appear to provide several advantages over<br />

previous methodologies. 35 PAMAM dendrimer nanoparticles covalently coupled with a fluorescent<br />

label can be visualized as sites of increased retention. 36 Any strategy to produce nanoparticles that<br />

combines diagnostic and therapeutic elements, however, must consider potential conflicting<br />

aspects. Introduction of some heterocyclic anti-cancer molecules may act to quench fluorescent<br />

properties and the capacity to detect fluorescent labels.<br />

Some nanoparticles are particularly promising for cancer diagnosis because of their exceptional<br />

properties of detection using current radiographic and magnetic methods. Some new materials<br />

being prepared as nanoparticles will provide the potential for visualization using novel imaging<br />

methods that may lead to greater selectivity of signal and reduce false positives. 37 For example,<br />

quantum dots associated with metastatic cancer cells can be visualized using fluorescence emissionscanning<br />

microscopy. 38 Similarly, lipid-encapsulated liquid perfluorocarbon contrast media at the<br />

site of a tumor can be detected <strong>by</strong> ultrasonic acoustic transmission. 39 It is also possible to functionalize<br />

materials such as quantum dots to incorporate materials that can be photo-activated to<br />

enhance their activity or detection. 23<br />

3.3 NANOPARTICLE TARGETING<br />

Nanoparticles can be designed in a variety of ways to achieve targeted delivery. Some targeting<br />

strategies rely upon inherent properties of the particle, in particular, its composition, size, and<br />

surface properties. Furthermore, the particle itself can either be the agent being delivered, or it<br />

can be prepared to carry a cargo for delivery. Cargo release from the nanoparticles can occur while<br />

the nanoparticle is still relatively intact or through its decomposition. A number of methods have<br />

been described to integrate and retain cargo components within nanoparticles and these, in general,<br />

match to chemical or physical characteristics of the cargo with those of the material used to<br />

generate the nanostructure. For example, positively charged cargo can be held within the nanoparticle<br />

through interactions with an internal network such as a polyanionic polymer that resembles<br />

the organization of secretory granules synthesized <strong>by</strong> cells. 40 Alternately, organized complexes<br />

akin to coacervates proposed to participate in cell structure evolution can be formed between cargo<br />

and particle matrix. 40 Therefore, for some cancer-targeting strategies, one should consider not only<br />

compatibility of the nanoparticle carrier with its cargo but also degradation events that might affect<br />

temporal aspects of particle stability and cargo release.<br />

Some nanoparticles can be designed or delivered in such a way as to produce a default targeting<br />

event; other nanoparticles must be decorated on their surface to produce a targeted structure.<br />

Topical application of a nanoparticle system at the target site may be all that is required for a<br />

successful outcome. Such a simple approach is not typically sufficient for effective targeting of<br />

many cancers. Successful targeting may require reduction of inherent targeting tendencies for the<br />

material(s) used to prepare the nanoparticle. Depending upon the physical and chemical nature of<br />

the nanoparticle and the mode of administration, there can also be complicating factors that affect<br />

the effectiveness of the targeting method. Inherent targeting and complicating factors associated<br />

with some nanoparticles used for a targeted delivery can impart safety issues that must also be<br />

considered. With such characteristics, it is easy to see why active targeting of nanoparticles to<br />

cancers can be both complicated <strong>by</strong> competing biological events as well as facilitated <strong>by</strong> these same<br />

properties. 41<br />

3.3.1 INHERENT TARGETING<br />

Nanotechnology for Cancer Therapy<br />

The RES is composed of a series of sentinel cells located in several highly perfused organs,<br />

including the liver and spleen. 27 Nanoparticles can be rapidly cleared from the blood if they are<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 25<br />

recognized <strong>by</strong> RES cells in a non-selective fashion, typically before achievement of effective<br />

targeting. 13,42 In some instances, this inherent targeting can provide a means to selectively delivery<br />

materials. 43 In most cases, it is possible to modify the physical and chemical characteristics of<br />

nanoparticles to reduce their default uptake <strong>by</strong> the RES. 44 Methods to avoid the RES will be<br />

addressed in depth in other chapters in this text. In general, these measures follow principles<br />

initially outlined in the development of stealth liposomes that provided a means of extending the<br />

circulating half-life of a nanoparticle. Although PEG molecules of various lengths coupled using<br />

various chemistries 45 are frequently used in this approach, heparan sulfate glycosaminoglycans<br />

(HSGs) have also been shown to provide a protective coating that reduces immune detection. 46<br />

Interestingly, HSGs might be shed at tumors <strong>by</strong> tumor-associated heparanase activity.<br />

Another inherent targeting aspect of nanoparticles relates to the nature of tumor-associated<br />

vasculature. In general, nanoparticles smaller than 20 nm have the ability to transit out of blood<br />

vessels. Solid tumors grow rapidly; tumor-associated endothelial cells are continually bathed <strong>by</strong> a<br />

plethora of cancer cell-secreted growth factors. In turn, endothelial cells sprout new vessels to<br />

provide needed nutrients for the continued growth of the tumor. This cancer cell-endothelial cell<br />

relationship, however, leads to the establishment of a poorly organized vasculature that, under the<br />

constant drive of growth factor stimulation, fails to organize into a mature vascular bed. Therefore,<br />

tumor-associated vascular beds are poorly organized and more leaky that normal vasculature.<br />

Nanoparticles will inherently target to tumors as exudates through leaky vasculature. This phenomenon,<br />

referred to as the enhanced permeability and retention (EPR) effect, 47 will be covered<br />

extensively in other chapters in this text.<br />

Finally, peculiar surface properties of certain nanomaterials might affect their inherent<br />

interactions that could act to detract from a targeted delivery strategy. For example, some polyanionic<br />

dendrimers can be taken up <strong>by</strong> cells and act within those cells to interfere with replication of<br />

human immunodeficiency virus (HIV) that is considered to be the causative agent of AIDS. 48<br />

Although, from such studies, it is unclear if these dendrimers interact with the host cell or the<br />

pathogen to block their interaction; such a finding points to the potential for nanoparticles to<br />

interact with structures that might affect their cellular properties or cell function. In some cases,<br />

a nanoparticle with inherent capacity to interact with a cell or tissue might provide an added<br />

advantage of using that material for a specific indication. In other cases, such an inherent capacity<br />

to bind to or recognize cell or tissue components might highlight potential distractive aspects of that<br />

material for certain indications.<br />

3.3.2 COMPLICATING ASPECTS<br />

By their eponymous descriptor, nanoparticles have physical dimensions in the nanometer size scale<br />

similar to viruses and other materials that are either recognized <strong>by</strong> the body as pathogens or are<br />

elements associated with an infective event. Toll-like receptors (TLR) present on monocytes,<br />

leukocytes, and dendritic cells play a critical role in innate immunity with the capacity to recognize<br />

organized patterns present on viruses and bacteria. 49 TLR proteins are present on the surface of<br />

cells in the lung, spleen, prostate, liver, and kidney. Because the patterned surfaces of nanoparticles<br />

can look like pathogen components recognized <strong>by</strong> TLR proteins such as DNA, RNA, and repeating<br />

proteins like flagellin, it is possible that a number of cell types might non-selectively interact with<br />

some nanoparticles. If such an interaction occurs, there are several potential outcomes that might<br />

produce complicating aspects for nanoparticle targeting to cancers. Nanoparticle materials might be<br />

immediately recognized and cleared <strong>by</strong> cells of the innate immune system, limiting the usefulness<br />

of even their initial application. Alternately, only a fraction of the applied nanoparticles might<br />

engage TLR that would not significantly affect the effectiveness of the administration. Recognition<br />

of even a small fraction of the administered nanoparticles, however, might lead to immune events<br />

that diminish the effectiveness of subsequent administrations. Nanoparticles that the naïve body<br />

initially tolerates may become a focus of immune responses upon repeated exposure.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


26<br />

Oppositely, one could envisage how potential recognition <strong>by</strong> TLR proteins might be beneficial<br />

in certain applications. In particular, cancer cells or tumors sites that express a particular TLR might<br />

actually provide a unique targeting aspect for nanoparticles prepared from the right material. In<br />

fact, some studies have been described using empty RNA virus capsules from cowpea mosaic virus<br />

as biological nanoparticles for delivery. 50 Because tumors can also contain a number of immune<br />

cells with some that might express TLR, the potential exists to have inherent targeting of nanoparticles<br />

to cancers through these tumor-associated cells. Such a circumstance could reduce the<br />

complexity of construction for a targeted nanoparticle complex. Such a suggestion has solid<br />

grounding in the extensive use of nanoparticles as adjuvants used for vaccination. 51 Additionally,<br />

nanoparticles such as liposomes can be selectively directed specifically to Langerhans cells in the<br />

skin as a means of enhancing the delivery of antigens to these professional antigen presenting<br />

cells. 52<br />

Recognition of nanoparticles <strong>by</strong> TLR proteins may or may not act to stimulate potential target<br />

cells. If stimulation occurs, the result can include release of cytokines and a variety of potent<br />

regulatory molecules. Induction of pro-inflammatory cytokines may provide an undesirable<br />

outcome because cancerous states appear to be motivated <strong>by</strong> inflammatory events. 53 The potential<br />

for a beneficial or negative impact of such outcomes would require case-<strong>by</strong>-case scrutiny. By<br />

themselves, nanoparticles have been shown to stimulate several cell-signaling pathways, including<br />

those that drive the release of pro-inflammatory cytokines such as interleukin (IL)-6 and IL-8. 54,55<br />

The incorporation of some therapeutic or diagnostic agents into nanoparticles can further enhance<br />

this potential for an inflammatory outcome. 56 Additionally, a variety of nanoparticles, including<br />

fullerenes and quantum dots, has been shown to stimulate the creation of reactive oxygen species,<br />

and this effect can be enhanced <strong>by</strong> exposure to UV light that might be used for visualization and/or<br />

activation (reviewed in Oberdorster, Oberdorster, and Oberdorster 57 ). As discussed at the beginning<br />

of this section, the potential for such events may be desirable for induction of selected events, e.g.,<br />

immunization against cancer cell antigens. Nanoparticles can also be prepared to release cytokines<br />

that might affect an anti-tumor immune events event. 58<br />

3.3.3 SAFETY ISSUES<br />

Nanotechnology for Cancer Therapy<br />

Although generalizations can be made for a particular targeted nanoparticle delivery system,<br />

specific issues will arise for each type of payload it contains and for each indication. It will be<br />

important to balance potential safety concerns for using nanoparticles with their potential benefits<br />

for reducing a safety concern that occurs without their use for comparable (or even improved)<br />

efficacy. Nanoparticles can tremendously reduce toxicity <strong>by</strong> sequestering cytotoxic materials from<br />

non-specific tissues and organs of the body until reaching the cancer site. 59 Polymeric nanoparticles<br />

have been loaded with tamoxifen for targeted delivery to breast cancer cells to improve the efficacy<br />

to safety quotient relative to direct administration of this cytotoxic agent. 60 Coupling of doxorubicin-loaded<br />

liposomes to antibodies or antibody fragments that can enhance targeting to cancer<br />

cells appears quite promising as a means of further improving the efficacy to toxicity profile for this<br />

chemotherapeutic. 61 PEGylated liposomal doxorubicin has improved tolerability with similar efficacy<br />

compared to free drug. 62 Liposomal formulations of anthracyclines appear to improve the<br />

cardiotoxicity profile of this anti-neoplastic. 63 PAMAM dendrimers conjugated to cisplatin not only<br />

improved the water solubility characteristics of this chemotherapeutic but also improve its toxicity<br />

profile. 64 PAMAM dendrimers have also been used to increase the effectiveness of a radioimmunotherapy<br />

approach <strong>by</strong> increasing the specific accumulation of radioactive atoms at a tumor site as<br />

well as improving selectivity of biodistribution. 65<br />

As previously mentioned, potential inherent targeting of some nanoparticles may initiate<br />

cellular responses, following recognition <strong>by</strong> immune cell surface receptor systems. Outcomes<br />

from such events could produce safety concerns, particularly for repeated exposures. Data<br />

suggest additional safety concerns related to the material(s) used in nanoparticle preparation as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 27<br />

well as nanoparticle size for some routes of administration. Recent studies have called into question<br />

the safety of repeated aerosolized exposure of carbon nanomaterials, 66 and nanoparticles have been<br />

shown to have a higher inflammatory potential per given mass than do larger particles. 57 Such<br />

particles may provide an efficient means to actively target lung cancers that is facilitated <strong>by</strong> the<br />

natural properties of these materials. In general, safety issues related to using nanoparticles to target<br />

cancers for treatment and diagnosis will be specific for the material used to prepare the particle: its<br />

size, the type of targeting mechanism it utilizes, its fate at the targeted site, and its non-targeted<br />

distribution and elimination pattern from the body.<br />

3.4 TARGETING TO CANCER CELLS<br />

In a very simplified sense, cancers occur through dysregulation of normal cell function. The body is<br />

designed to correct tissue defects following damage, to expand selected immune cells in response to<br />

a pathogen, and to compensate for perceived deficiencies in one cell type <strong>by</strong> altering the phenotype<br />

of another such as in stem cell recruitment. Each of these processes is mediated <strong>by</strong> (normally)<br />

regulated events that allow cells to lose their differentiated phenotype with restrained growth<br />

characteristics and acquire a replication-driven phenotype. A lack of re-differentiation into a<br />

growth-restrained, differentiated phenotype is the paradigm of cancer. Repair of an epithelial<br />

wound is a good example of this phenomenon (Figure 3.3). It is the lack of re-differentiation<br />

and continued responsiveness of these cells to growth factors and growth activators that supports<br />

and maintains the cancer phenotype. Extensive genomic differences between differentiated and<br />

non-differentiated forms of the same cell account for the differences observed between these two<br />

cell phenotypes. 67<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

Growth<br />

factors<br />

Cell division<br />

FIGURE 3.3 Loss and recovery of barrier function associated with normal epithelia. (a) Epithelia express tight<br />

junctions ( ) and associate at their base with a complex of proteins known as the basement membrane<br />

( ). (b) Damage to epithelia increases responses to growth factors and results in differentiated epithelial<br />

cells that can freely divide. (c) Cell division continues until damaged area is covered. (d) Cell–cell contacts<br />

such as adherens junctions, tight junctions, and gap junctions have been shown to suppress growth and<br />

stimulate re-differentiation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


28<br />

Many of the differences between replicating and non-replicating cells are associated with<br />

surface and metabolic properties that can be used to discriminate between differentiated<br />

(normal) and de-differentiated (cancer) cells. One of the biggest concerns using this information<br />

to target cancer cells is that non-cancerous cells undergoing normal and necessary repair processes<br />

may transiently express these same targets. For example, herceptin is an antibody that binds to<br />

her2/neu receptors that are over expressed on the surfaces of cancer cells. Unfortunately, this<br />

antibody can also target normal cardiac cells undergoing repair induced <strong>by</strong> the actions of a<br />

common anti-cancer agent, doxorubicin; patients on doxorubicin treatment are placed on an<br />

extended washout period prior to exposure to herceptin. Therefore, nanoparticles having a cytotoxic<br />

capacity and targeted using the herceptin antibody could result in cardiomyopathy. Fortunately, the<br />

high, transient, systemic levels of doxorubicin associated with direct administration can be muted<br />

<strong>by</strong> administration in nanomaterials such as liposomes, reducing the risk of cardiomyopathy. 68 That<br />

many surface and metabolic properties are the same for both cancer cells of a particular cell type<br />

and the undifferentiated form of that cell type during normal cell function must be appreciated as<br />

one examines potential cancer cell-targeting strategies for nanoparticles.<br />

The general issues raised above concerning safety aspects of targeting cancer cells highlight<br />

concerns of selecting a strategy that properly accounts for unique cell surface properties and<br />

metabolic activities of cancer cells relative to normal cells. All too frequently, normal cells can<br />

undergo processes (e.g., wound repair) that will transiently transform them into a cell with surface<br />

properties or metabolic characteristics indistinguishable from a cancer cell. In some aspects, these<br />

altered surface properties and metabolic activities are intertwined. Alterations in metabolic properties<br />

can induce increased expression of nutrient uptake pathways and catabolic proteins that assist<br />

in nutrient absorption to sustain the accelerated growth rate of cancer cells. Because some of these<br />

components are present at the cell surface, a cancer cell’s composition and profile are modified <strong>by</strong><br />

their presence. From an opposing perspective, increased surface expression of components such as<br />

growth factor receptors will shift the metabolic activity of a cancer cell following activation of that<br />

receptor (either constitutive activation or ligand-induced activation).<br />

3.4.1 CELL SURFACE PROPERTIES<br />

Nanotechnology for Cancer Therapy<br />

In general, one thinks of targeting cancer cells <strong>by</strong> use of a highly specific surface material that<br />

absolutely identifies only that cancer cell within the entire body. Some studies, however, have<br />

demonstrated that cancer cells growing in different sites of the body can have altered surface<br />

properties that could facilitate non-specific nanoparticle binding and uptake; colloidal iron hydroxide<br />

(CIH) nanoparticle association and uptake appears to be enhanced for transformed cells, 69 and<br />

CIH nanoparticles show differences in cell surface interactions following transformation of chick<br />

embryo fibroblasts. 70 Such observations suggest that generic cell-surface charge differences might<br />

provide a targeting strategy for nanoparticles that could be, even if not highly discriminating, used<br />

to enrich nanoparticle delivery to cancer cells through non-specific associations. In combination<br />

with other targeting strategies, surface charge differences might provide a useful adjunct. For<br />

example, some cancer cells express unique sets of surface enzymes that might be useful to activate<br />

a prodrug once a nanoparticle has been localized to the surface of a cancer cell. In this regard, a<br />

number of proteases have been shown to be significantly up-regulated <strong>by</strong> oncogenic conversion. 71<br />

A proof of concept for this type of specific application has been described using a polymer-based<br />

fluorogenic substrate PB-M7VIS that serves as a selective proteobeacon. 36<br />

A number of overexpressed growth factor receptors have been used to selectively target cancer<br />

cell surfaces, and the description of many of these targets has been reviewed. 72 With regard to<br />

targeting nanoparticles, it is important to remember that once engaged <strong>by</strong> the targeting ligand, some<br />

of these surface components are internalized whereas others will remain at the cell surface.<br />

Matching the type of nanoparticle material and its potential cargo with the likely fate of the targeted<br />

structure can be critical to optimizing the desired outcome. It is also possible that once bound,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 29<br />

ligand–nanoparticle complexes could lead to receptor internalization that might not occur <strong>by</strong> the<br />

presence of the targeting ligand alone. The basis for this difference might come from the potential<br />

for nanoparticles to contain a coordinated ligand matrix to sequester of cell-surface receptors in a<br />

manner that facilitates internalization.<br />

Antibody-based targeting of nanoparticles to solid tumors has been a highly promising strategy<br />

that is augmented <strong>by</strong> the enhanced vascular permeability (EPR effect) of solid tumors. For example,<br />

an antibody-directed (anti-p185HER2) liposome loaded with an anti-neoplastic can be an effective<br />

cancer therapeutic approach. 73 Blood cell-based cancers (e.g., leukemias and lymphomas) can also<br />

be targeted <strong>by</strong> nanoparticles as a way to reduce unwanted systemic side effects. Nanoparticles could<br />

be targeted to T-cell leukemia cells using an antibody to a surface cluster of differentiation (CD)<br />

antigen, CD3, on the surface of lymphocytes. 74 B-cell lymphomas can be targeted <strong>by</strong> anti-idiotypic<br />

antibodies specific for the unique monoclonal antibody expressed <strong>by</strong> each individual cancer. 75 In<br />

both of these cases, the potential for these B- and T-cell-derived cancer cells to actively take up<br />

particles on their surfaces as part of their normal function in antigen surveillance and presentation<br />

might facilitate and even augment the desired outcome using a targeted nanoparticle.<br />

Efficient, targeted delivery of gene therapy elements and/or antigens to antigen presentation<br />

cells (APC) has long been a goal for the induction of anti-cancer cell immune responses. Nanoparticles<br />

provide an exciting possibility to achieve this goal. Coating nanoparticles with mannan<br />

facilitates their uptake <strong>by</strong> APCs such as macrophages and dendritic cells that acts to target these<br />

materials to local-draining lymph nodes following their administration. 76 Such an approach is likely<br />

to provide additional synergy in APC activation because polymer nanoparticles are efficiently<br />

phagocytosed <strong>by</strong> dendritic cells. 77 Many APCs express LDL-type receptors, and molecules that<br />

interact with this class of receptors could be a means of targeting as well. 78 Interestingly, LDL<br />

receptors can be an attractive targeting strategy for cancers because many tumors of different<br />

origins express elevated levels of this receptor. 79 Therefore, LDL-based nanoparticles could be<br />

useful in targeting cancer. 31<br />

It might also be possible to intentionally alter the surfaces of cancer cells to improve nanoparticle<br />

targeting. Cells could be transfected with a protein that expresses the appropriate acceptor<br />

peptide recognized <strong>by</strong> a surface-applied bacterial biotin ligase. 80 Although there would be multiple<br />

issues to overcome prior to clinical application, this approach outlines one strategy where cancer<br />

cells might be altered to express a unique surface structure such as biotin that could be very<br />

selectively targeted. Reversed-response targeting might also be performed using discriminating<br />

cell-surface properties. Hepatocytes can be targeted using nanocapsules decorated with the surface<br />

antigen of hepatitis B virus (SAgHBV). 81 Because liver cells may lose their capacity to bind<br />

SAgHBV following oncogenic conversion, this targeting strategy could be used to deliver cytoprotective<br />

materials to normal hepatocytes and enhance the efficacy of chemotherapeutics aimed<br />

at liver cancers. Similarly, hepatocytes exclusively express high affinity cell-surface receptors for<br />

asialoglycoproteins, and this ligand–receptor system has been used to target albumin nanoparticles<br />

to non-cancer cells of the liver. 8 Nanoparticles coated with galactose might also be used to target<br />

the liver. 82<br />

3.4.2 METABOLIC PROPERTIES<br />

One of the most detrimental aspects of cancer cells, their high rate of proliferation, can also be<br />

considered their Achilles’ heel. As proliferation rate increases, metabolic requirements follow<br />

accordingly. This places cancer cells in a precarious position where a blockade of critical metabolic<br />

steps can lead to cytotoxic outcomes; this is the basis from a number of currently approved anticancer<br />

agents that function as metabolic poisons. 83 There are two obvious approaches that could be<br />

used for targeting using this characteristic of cancer cells: ligands that emulate the nutrient, vitamins<br />

or co-factors, and antibodies that recognize these surface transport elements. Nanoparticles<br />

coated with a ligand for one of these receptors such as folate can be used to target to cancer cells. 84<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


30<br />

Folate has been used to target dendrimers 85,86 and iron oxide nanoparticles. 87 Folate receptortargeted<br />

lipid nanoparticles for the delivery of a lipophilic paclitaxel prodrug have shown promising<br />

pre-clinical outcomes. 88 Cancer cells can also overexpress transferrin receptors (REF), and transferrin-conjugated<br />

gold nanoparticle uptake <strong>by</strong> cells has been demonstrated. 89 A transferrinmodified<br />

cyclodextrin polymeric nanoparticle has been described that could be used to deliver<br />

genetic material to cancer cells. 90<br />

Increased nutrient uptake associated with increased requirements for amino acids and nucleic<br />

acids may also provide a targeting strategy for nanoparticles. That many classical anti-cancer agents<br />

function <strong>by</strong> interfering with amino acid or nucleic acid incorporation into polymer structures should<br />

provide an important template for strategies to use nanoparticle technologies to effectively target<br />

cancer cells. For example, pancreatic cancer cell lines appear to overexpress the peptide transporter<br />

system PepT1, 91 possibly providing a growth advantage <strong>by</strong> its ability to provide additional amino<br />

acid uptake. Nanoparticles decorated with (or composed of) ligands recognized <strong>by</strong> this uptake<br />

pathway may provide an important targeting opportunity. Such an approach can be used for<br />

similar target-specific delivery of other molecules. It is important to remember that materials such<br />

as amino acids and nucleic acids, unlike co-factors discussed above that are used more as part of<br />

catalytic cellular events, are required in stoiciometric amounts for cell growth. Targeting strategies<br />

using uptake processes against vitamins and co-factors will have the added benefit of potentially<br />

depriving cancer cells of these critical materials whereas targeting strategies using amino acids and<br />

nucleic acids will probably not affect the overall influx of these materials and their incorporation into<br />

nascent polymers required for continued cell growth.<br />

A growing bank of experimental and clinical data has provided strong evidence that chronic<br />

inflammation can drive epithelial cell populations into an oncogenic phenotype. 92 It is this preneoplastic<br />

character that may act to alter the metabolic character of cells that might be useful for<br />

targeting nanoparticles. Such a targeting strategy would make use of transitions in cellular function<br />

in response to pro-inflammatory signals (Figure 3.4). In this regard, one could envisage liganddirected<br />

targeting to inflammatory sites as well as activation of nanoparticles (or their components)<br />

for localized delivery at these sites <strong>by</strong> the presence of unique enzymatic activities. Additionally,<br />

one could contemplate nanoparticles that deliver cancer prevention agents that work through<br />

suppression of inflammatory events. A ligand peptide that binds endothelial vascular adhesion<br />

molecule-1 (VCAM-1) on the surfaces of inflamed vessels has been used to target nanoparticles. 93<br />

One major concern with using such metabolic processes for targeting nanoparticles is that inflammatory<br />

events are a common and essential function of the body, and they are not necessarily<br />

associated with pre-cancerous or cancerous states. Some nanoparticles can induce an inflammatory<br />

response. Therefore, the method selected for such a nanoparticle-targeting strategy most keep this<br />

in mind. It might be possible to utilize additional cancer-targeting mechanisms (e.g., the EPR<br />

effect) to augment inflammation-based strategies.<br />

3.5 TARGETING TO TUMORS<br />

Nanoparticle targeting strategies involving peculiar and unique properties of tumors have been<br />

described. Two of these properties will be discussed: aberrant vasculature and the unique metabolic<br />

environment produced <strong>by</strong> inefficient blood supply resulting from the aberrant vasculature. These<br />

targeting strategies have strong similarities to those described for targeting cancer cells, taking<br />

advantage of unique endothelial cell-surface properties in tumors or the cells’ peculiar environment<br />

as they respond to the metabolically overactive environment of a tumor.<br />

3.5.1 ABERRANT VASCULATURE<br />

Nanotechnology for Cancer Therapy<br />

Tumor neovasculature is a promising site for targeting nanoparticles for both diagnosis and therapy.<br />

Once a tumor reaches a size of greater than w1 cm 3 , vascular assistance to deliver oxygen and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 31<br />

(a)<br />

(b)<br />

(c)<br />

Macropage<br />

Macropage<br />

Cell division<br />

Macropage<br />

FIGURE 3.4 Effect of chronic pro-inflammatory stimulus on epithelial barrier patency. (a) A variety of stimuli<br />

( ) can incite the release of pro-inflammatory cytokine (e.g., TNFa and IFNg) from macrophages and other<br />

cells. Genetic predisposition can enhance these responses. (b) Released pro-inflammatory cytokines act to open<br />

tight junction (TJ) structures ( ), allowing entry of additional activating stimuli that, in turn, attract more cells<br />

associated with inflammation. (c) Chronic inflammation leads to breakdown of basement membrane, loss of TJ<br />

function, and disorganization of the epithelia characteristic with a pre-neoplastic state.<br />

remove <strong>by</strong>-products is required for cell survival. Cancer cells secrete a variety of growth factors<br />

that stimulate the formation of nascent blood and lymph vessels. Tumor-associated vascular beds<br />

have unique surface properties that are associated with their rapid growth characteristics, resulting<br />

in aberrant vascular beds that have been used to design a combined vascular imaging and therapy<br />

approach using nanoparticles. 94 Non-specific targeting of nanoparticles (10–500 nm in diameter) to<br />

solid tumors through this EPR capacity of solid tumors 47 can provide a means to enrich the<br />

localization of an anti-neoplastic agent to a tumor when it is coupled to a nanoparticle compared<br />

to its free form. 95 The EPR effect has also been used to increase localization through inherent<br />

targeting 96 of long-circulating liposomes. 97<br />

PAMAM dendrimers useful for boron neutron capture therapy (BNCT) of cancers have been<br />

targeted to tumor vasculature <strong>by</strong> attachment of vascular endothelial growth factor (VEGF) that acts<br />

to target VEGF receptors that are frequently overexpressed on tumor neovasculature. 98 Targeting<br />

VEGF receptors Flk or Flt on tumor-associated endothelial cells could also be effective; this has<br />

been done with a complex material composed of anti-Flk-1 antibody-coated 90 Y-labeled nanoparticles<br />

99 . PECAM (or CD31) is highly expressed on the surface of endothelial cells present in<br />

immature vasculature. Platelet endothelial cell adhesion molecule (PECAM) up-regulation<br />

occurs following VEGF stimulation of endothelial cells. The presence of such endothelial<br />

surface markers also provides the opportunity to target nanoparticles that contain DNA for gene<br />

therapy applications, 100 release anti-angiogenesis agents as well as chemotherapeutics, 101 and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


32<br />

antigens/agents to stimulate an anti-cancer cell immune response. 101,102 Vascular cell adhesion<br />

molecule-1 (VCAM-1) is a marker for inflammation of the endothelial and has been used to target<br />

nanoparticles to these sites for magento-optical imaging. 103<br />

Differences in tissue-specific endothelial surface properties might be used for targeting nanoparticles.<br />

104 In one such example, a peptide that binds to membrane dipeptidase on lung endothelial<br />

cells can be used to target nanocrystals to the lung. 105 Doxorubicin-loaded nanoparticles have been<br />

targeted to tumor vasculature <strong>by</strong> surface decoration with a cyclic arginine–glycine–aspartic acid<br />

(RGD) peptide that binds to the cell adhesion molecule integrin avb3 on the surface of endothelial<br />

cells. 106 A peptide that interacts with the lymphatic vessel marker podoplanin can be used to target<br />

nanocrystals to lymph vessels and some tumor cells. 105 Cationic nanoparticles, containing genes<br />

that can block endothelial cell signaling that were selectively targeted to tumor vasculature <strong>by</strong><br />

couplingtoanintegrina vb 3 ligand, were shown to produce endothelial apoptosis and tumor<br />

regression. 107<br />

3.5.2 METABOLIC ENVIRONMENT<br />

Solid tumors typically grow at such a rapid pace that vascular function fails to keep pace with local<br />

metabolic requirements. This situation, leading to reduced oxygen tension and a depressed pH as a<br />

result of a lack of waste acid clearance, is associated with the onset of necrotic cores of large<br />

tumors. A number of studies have shown that cancer cells become adapted to successful growth in<br />

these difficult conditions, providing unique characteristics that might be exploited for tumor<br />

targeting. For example, nanoparticles will accumulate at inflammation sites along the bowel<br />

when administered orally. 108,109 Because these inflammatory sites are typically associated with a<br />

slightly acidic environment, a similar targeting strategy may be possible for tumors where the pH<br />

has also been depressed. Nanoparticles targeted to cancer cells might also be modified at low<br />

oxygen tension to either release or activate a payload. Frequently, cancer cells rely more on<br />

glycolysis than oxidative metabolism for energy production. This finding is consistent with the<br />

reduced oxygen tension of poorly perfused tumors. These conditions would result in the release of<br />

acidic glycolytic end-products that are not effectively cleared <strong>by</strong> the sluggish blood flow through<br />

tumor vascular beds. Therefore, targeting that takes advantage of a slightly depressed local pH that<br />

might be found in some tumors could be an attractive design strategy.<br />

3.6 FATE OF NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

Successful targeting of nanoparticles to cancers or tumors may involve overcoming multiple<br />

biological, physiological, and physical barriers. Site of initial application can be critical. For<br />

example, nanoparticles administered into the gut or lung would initially confront epithelial barriers.<br />

Metabolic events or cellular responses at an injection site represent another initial barrier to targeted<br />

nanoparticle delivery. Once nanoparticles have entered the body, their size, shape, or surface<br />

characteristics can initiate events that present a second barrier to targeted delivery—misdirection<br />

of the material away from its targeted site through undesired interactions. Access and/or enriched<br />

distribution to specific organs or regions of the body may be critical for successful nanoparticle<br />

targeting. Finally, once nanoparticles have reached a targeted site, metabolic or physical aspects of<br />

the cancer cell or tumor might limit their effectiveness. Here, the intracellular uptake and fate may<br />

dictate the potential success of each nanoparticle approach. Events at each of these barriers act in a<br />

cumulative fashion to limit the success of any nanoparticle-based targeting strategy. Obviously, the<br />

overall fate of nanoparticles might be improved <strong>by</strong> using materials that are not affected <strong>by</strong> these<br />

barriers or <strong>by</strong> modification of nanoparticles in ways that can neutralize these barrier issues.<br />

Matching the size and composition profile of a nanoparticle delivery system with the targeting<br />

strategy allows for an optimized approach that takes advantage of default targeting events as much<br />

as possible.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 33<br />

As previously discussed, it is critical that inherent targeting mechanisms associated with a<br />

particular nanoparticle does not overwhelm or work in concert with any applied targeting strategy.<br />

Once at the target site, the fate of the nanoparticle can significantly affect its potential to provide the<br />

desired outcome. The delivery of hybridization-competent antisense oligonucleotides (ODNs)<br />

targeted to a cancer cell or tumor would not provide the desired outcome unless this material is<br />

efficiently internalized. ODNs covalently conjugated to anionic dendrimers have been shown to<br />

effectively deliver through an endocytosis process and down-regulate epidermal growth factor<br />

receptor expression in cancer cells. 110 Other targeting strategies may not provide the desired<br />

outcome if the nanoparticle is internalized <strong>by</strong> cells. For example, enzyme-coupled nanoparticles<br />

targeted to a tumor that would activate a prodrug could fail to provide a desired outcome. Therefore,<br />

the potential for nanoparticles to have a successful outcome of targeting cancer cells or tumors<br />

requires favorable events at natural barriers of the body and their distribution within the body, but<br />

also their fate at the targeted site.<br />

3.6.1 SITE OF INITIAL APPLICATION<br />

Several extracellular barriers exist for the administration and targeted delivery of nanoparticles.<br />

Initial entry into the body represent on obvious barrier. This is not an issue for situations where the<br />

cancer cell or tumor targeted is readily accessible <strong>by</strong> a topical application. Such a situation,<br />

however, is rather rare. Entry into the body across mucosal surfaces such as those in the gut or<br />

lung is typically very inefficient. Even viral particles are not very successful at this approach with<br />

most relying upon infecting cells of the barrier from the apical exposure <strong>by</strong> only a few viral particles<br />

that can replicate inside the cells to allow the basolateral (systemic) release of large numbers of<br />

progeny. Viral particle entry at apical surfaces of epithelial cells is decreased <strong>by</strong> physical barriers<br />

such as secreted mucus as well as proteases and other enzymatic barriers. Extracellular (acellular)<br />

matrix environments that viruses might encounter after systemic infection could similarly act to<br />

diminish cellular targeting and entry. Man-made nanoparticle delivery systems are likely to be<br />

impeded <strong>by</strong> these same physical and biological barriers at epithelial surfaces and within the body.<br />

Reduced surface exposure of highly charged or protruding structures is commonly used <strong>by</strong> viruses<br />

to minimize the impact of these extracellular barriers on viral infectivity. Similar considerations<br />

may facilitate optimization of nanoparticle delivery strategies.<br />

There are several common methods for administering materials to the body: injection or application<br />

to an epithelial surface (skin, intestine, lung, etc.) of the body. Nanoparticles can be absorbed<br />

into the skin after topical application. 111 Although nanoparticles can be taken up through appendages<br />

of the skin (sweat gland ducts, hair follicles) following topical application, 112 microneedles can be<br />

used to dramatically increase the efficiency of their uptake into and across skin. 113 The intravitreous<br />

injection of nanoparticles results in transretinal movement with a preferential localization in retinal<br />

pigment epithelial cells, 114 allowing for a sustained delivery strategy to the inner eye.<br />

Nanoparticles can be absorbed from the lumen of the gut, but this absorption is inefficient. 115 A<br />

number of factors have been examined related to regulation of nanoparticle uptake from the gut<br />

lumen. 116 Nanometer-sized liposomes enter into the intestinal mucosa better than larger, multilamellar<br />

liposomes, and this uptake can be improved <strong>by</strong> coating with a mucoadhesive polymer such<br />

as chitosan. 117 It is interesting that lipid-based materials absorbed from the gut partition into the<br />

lymphatic system and studies have suggested that these particles have remarkable access to the<br />

hepatocytes. 118 One way to potentially improve nanoparticle uptake from the gut is to PEGylate<br />

these materials in a manner that selectively increases binding to the intestinal mucosa rather than<br />

the stomach wall. 119 Additionally, anionic PAMAM dendrimers have been shown to rapidly cross<br />

the intestinal mucosa in vitro and may provide a method to improving oral delivery of nanoparticles.<br />

120 Cationic dendrimers also show a transcytosis capability in vitro; in general, cationic<br />

dendrimers are more cytotoxic than anionic dendrimers, but this characteristic can be reduced <strong>by</strong><br />

additional surface modifications using lipids. 121 Formulation studies have been performed to<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


34<br />

identify optimal methods for aerosol delivery of nanoparticles to the lung. 122 In general, the uptake<br />

of nanoparticles at the lung or gut surface occurs, but the efficiency of this uptake is dramatically<br />

improved <strong>by</strong> incorporation of a specific uptake mechanism. Even without a specific uptake<br />

mechanism, an appreciable amount of nanoparticle absorption can occur at these sites if they are<br />

sufficiently stable and are not removed <strong>by</strong> clearance mechanisms.<br />

A large number of studies have been performed to assess nanoparticle absorption following<br />

inhalation exposure, and concerns over the safety of such an approach for drug delivery have been<br />

raised. 123 Nanoparticles deposited in the airways appear to be taken up through transcytosis pathways<br />

that allow the passage of these materials across epithelial and endothelial cells to reach the blood and<br />

lymphatics. 57 Surface properties of nanoparticles greatly affect the capacity of nanoparticles to be taken<br />

into cells through the process of endocytosis and uptake following pulmonary deposition. As part of the<br />

respiratory tree, intranasal administration of nanoparticles can potentially provide a route into the brain.<br />

3.6.2 DISTRIBUTION TO SPECIFIC ORGANS<br />

One of the most difficult challenges of administering cytotoxic chemotherapeutics involves<br />

unwanted exposure to non-cancer cells and to tissue and organ compartments not involved with<br />

the disease. Nanoparticles carrying a chemotherapeutic can reduce the undesirable distribution of<br />

such compounds as they are restricted from some compartments of the body such as the brain. 59<br />

Oppositely, nanoparticles can be modified, e.g., conjugation to chelators, to acquire the capacity to<br />

transport across the blood–brain barrier or BBB. 124 Alternately, nanoparticles coupled to certain<br />

protein ligands such as apolipoprotein E can be used to target and transport across the BBB. 125 In<br />

both cases, these nanoparticle delivery approaches lead to the unique distributions of materials that<br />

must be cleared and/or metabolized. Therefore, one consequence of targeting cancers cells is that<br />

the fate of these materials, <strong>by</strong> accessing and localizing to sites where cancer cells reside, may be<br />

affected that could affect their overall safety as well as efficacy.<br />

Targeting to some cancers may require overcoming additional hurdles beyond interaction with<br />

specific cancer cell or tumor components. Some tumors are located in difficult-to-reach sites such as<br />

the brain and testes. Accessing these sites from the systemic vasculature requires that nanoparticle<br />

materials must first avoid systemic clearance <strong>by</strong> the RES and have the capacity to move across<br />

either the blood–brain or blood–testes barrier. Whereas some types of nanoparticles can keep<br />

materials out of compartments of the body such as the brain, 59 other types of nanoparticles may<br />

provide access to this difficult-to-reach compartment. 126 In general, surface characteristics that can<br />

be altered through chemical modifications can be used to regulate the targeted delivery of nanoparticles<br />

to specific sites within the body such as the brain (reviewed in Olivier 14 ). It has even been<br />

reported that coating nanoparticles with polysorbate 80 can facilitate brain targeting <strong>by</strong> enhancing<br />

their interaction with brain microvasculature. 127 Alternately, intranasal delivery of macromolecules<br />

has been suggested to move in a retrograde fashion into the brain following uptake at the olfactory<br />

epithelium. 128 Polylactic acid–PEG nanoparticles have been shown to transport across the nasal<br />

mucosa 129 and could, theoretically, also provide some access to the brain because both materials<br />

appear to transport via a transcytosis mechanism.<br />

3.6.3 INTRACELLULAR UPTAKE AND FATE<br />

Nanotechnology for Cancer Therapy<br />

Particularly in the case of cancer therapeutics, some targeted delivery may require access to intracellular<br />

sites such as the nucleus or mitochondria. In general, nanoparticles with diameters less than<br />

50 nm can easily enter most cells. Successful cellular uptake of nanoparticle systems targeted to cancer<br />

cells and/or tumors, however, frequently depends upon the balance of mechanisms that act to clear<br />

nanoparticles from the circulation and mechanisms that allow for their retention in this compartment.<br />

Several clearance mechanisms exist, and loss of nanoparticles from the circulation appears to be<br />

dominated <strong>by</strong> macrophage uptake, following complement activation or surface opsinization. 130<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 35<br />

Some agents such as commonly used chemotherapeutics are capable of moving efficiently across the<br />

plasma membrane of cells and into the cytoplasm that allows access to intracellular organelles if the<br />

material can be released locally to the target site. A nanoparticle-based delivery may require the release<br />

of a cargo either at the cell surface or after internalization of the nanoparticle <strong>by</strong> the target cell. Uptake of<br />

nanoparticles into cells can occur through a clathrin-coated endocytosis event, 131 through caveolae<br />

structures, 132 or through uptake mechanisms that do not appear to involve clathrin or caveolae. 133<br />

Differences in surface properties and nanoparticle size will likely dictate the predominant route of entry<br />

into a cell. Such uptake mechanisms into cells following nanoparticle targeting to a cancer cell or tumor<br />

can involve vesicular trafficking to acidified, protease, and nuclease-enriched lysosomal compartments<br />

within the cell.<br />

Unless the nanoparticle carrying a chemotherapeutic agent can release it prior to the trafficking<br />

of these vesicles to destructive (lysosomal) pathways or it can avoid such a pathway once inside the<br />

cell, the effectiveness of the absorbed material may be dramatically reduced. Also, some new<br />

classes of anti-cancer agents have poor membrane permeability properties and would not readily<br />

leave the endosome after uptake. Furthermore, exposure of these materials to lysosomal environments<br />

would destroy their biological activity; nucleic acid- or peptide/protein-based therapeutics<br />

capable of marking a cancer cell for clearance <strong>by</strong> the immune system would be examples of this<br />

type of approach. In these cases, proper selection of the nanoparticle composition and characteristics<br />

allows these materials to escape the fate of this default uptake event. There are endogenous<br />

properties of some materials as well as the capacity to include specific intracellular targeting agents<br />

that can be matched with the intracellular delivery desired for the material being targeted.<br />

Following specific (or non-specific) targeting of nanoparticles to a cell, a number of events act<br />

to traffic these material within the cell. In general, nanoparticles that have associated with a specific<br />

cell-surface target are internalized through an endocytosis process that produces the formation of<br />

intracellular vesicles, containing the nanoparticle. Based upon their physical and chemical characteristics,<br />

most nanoparticles reaching these endosomal vesicles are likely delivered to lysosomes<br />

within cells where they would be metabolized or retained. Therefore, nanoparticles would not<br />

readily access the cytoplasm of target cells. An ultimate fate of lysosomal structures within targeted<br />

cells is not necessarily a problem. Many diagnostics have already performed their function and/or<br />

can still be detected within this compartment. In the case of therapeutics, many of these may have<br />

already been released from the nanoparticle prior to its arrival at the lysosome, and/or the materials<br />

are stable in this hostile environment and continue to act upon the target cell from this location.<br />

However, there are a number of potential therapeutic compounds that will be inactivated <strong>by</strong> this<br />

outcome and that require additional delivery events to achieve their optimal function on the target<br />

cell. Oligonucleotide delivery to tumors is one such example where the therapeutic must access the<br />

target cell cytoplasm for its desired action. 134<br />

Reduction in the rate of endocytosis of nanoparticles can be achieved <strong>by</strong> coating them with<br />

proteins such as lactoferrin or ceruloplasmin that act to retain the material at the cell surface. 135 One<br />

approach to facilitate nanoparticle delivery to the cytoplasm of a target cell is to covalently attach<br />

the membrane penetrating TAt peptide derived from HIV-1 to the surface of these nanoparticles. 136<br />

Because nanoparticles of several compositions have been shown to target to the mitochondria,<br />

137,138 these materials may access the cell’s cytoplasm to reach this organelle. Such an<br />

outcome may be driven <strong>by</strong> the physico-chemical characteristics of the nanoparticle with relation to<br />

the unique proton and ion gradients found in mitochondria. Modification of the nanoparticle to<br />

affect this inherent targeting may be important.<br />

3.7 CONCLUSIONS<br />

Nanoparticles provide a range of new opportunities to increase the targeting of currently approved<br />

diagnostic and therapeutic agents to cancers. Improvements in targeting can lead not only to<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


36<br />

Nanotechnology for Cancer Therapy<br />

increased efficiency of these agents but also to increased signal-to-noise ratios for diagnostics and<br />

better efficacy to toxicity ratios for therapeutics. Currently, a whole new spectrum of biopharmaceuticals<br />

and biotechnological agents for cancer diagnosis and therapy are also being developed.<br />

Some of these materials require special formulation technologies to overcome drug-associated<br />

problems. Although nanoparticles offer improved profiles for some currently approved diagnostic<br />

and therapeutic agents, many biotechnology-based materials absolutely require some method of<br />

delivery that compensates for their poor stability or non-selective activity in a systemic setting.<br />

Nanoparticles offer a set of new opportunities for the development of these agents. 29<br />

Nanoparticles can be prepared in such a way as to have diagnostic or therapeutic agents<br />

integrated into them in ways that either freely releases the agents or that requires decomposition<br />

of the nanoparticle for the release to occur. Because of the inherent nature of small (nanometersized)<br />

structures, the body can identify and respond to these as foreign. Such a response can <strong>by</strong><br />

suppressed <strong>by</strong> incorporation of agents that might suppress undesirable responses, or the application<br />

can be matched to the nanoparticle to make use of these natural responses. It is even possible to<br />

modify these natural responses to better match the desired clinical outcome. Specific components<br />

used to prepare nanoparticles can affect not only their stability in the body but also their capacity to<br />

be absorbed across natural barriers of the body (e.g., BBB) as well as the inherent systemic<br />

distribution of the nanoparticle that might compete with or complement efforts to selective<br />

targeting strategies.<br />

Without the current fanfare related to nanotechnologies, nanoparticles have been used to<br />

selectively target a number of organs of the body for a number of years. Nanoparticle colloids<br />

were shown to have contrast media properties that related to the unique surface properties of cells in<br />

specific organs of the body. 139 Deviations from normal function such as oncogenic transformation<br />

can lead to changes in a cell’s surface properties and its capacity to interact with nanomaterials. For<br />

example, gadolinium-based nanoparticles are taken up <strong>by</strong> hepatocytes, and <strong>by</strong> the decreased function<br />

and density of cancer cells in the liver tumors, a reduced level of uptake of these particles can<br />

be used to identify tumors using T1-weighted images obtained from MRI. 42<br />

Throughout this chapter, there has been frequent referral to viral infection events as a paradigm<br />

for cellular and intracellular targeting strategies for nanoparticles. Indeed, these materials are very<br />

successful models for nanoparticle targeting because they have developed mechanisms to discriminate<br />

between the various cells of the body (e.g., tropism for only cells of the intestinal tract) and can<br />

deliver labile (polynucleic acids) payloads that dramatically affect cell function. In response to<br />

these nanoparticle invaders, host cells have established intricate and complicated mechanisms that<br />

block viral infectivity and cellular actions. Such evolutionary pressures have led to the incorporation<br />

of intricate and novel methods <strong>by</strong> viruses to effectively combat these protective systems<br />

established <strong>by</strong> host cells. It is into this environment where the virus nanoparticles and host cells<br />

have battled back and forth for millennia that efforts to use nanoparticles to deliver agents to<br />

cancers for diagnosis and/or therapy must be framed.<br />

Finally, it is important to sound a precautionary note for the potential to over-engineer nanoparticles.<br />

Nanoparticles provide a platform that can potentially be used to simultaneously function<br />

in targeting therapeutic molecules as well a reporter and/or imaging agents. Elegant studies have<br />

been performed with nanoparticles modified three, four, or even five times with materials that<br />

promoted active targeting and/or reduced non-specific targeting as well as corrected undesirable<br />

properties of residence, biodistribution, and stability. Such tour-de-force efforts would be<br />

considered unrealistic <strong>by</strong> pharmaceutical companies for scaling to a process that would gain<br />

approval from regulatory agencies. Clinical success can be demonstrated for a nanoparticle<br />

system only if it can get to the clinic; a viable production process is critical to this development<br />

path. Therefore, successful applications of nanoparticles in target cancers for therapy and diagnosis<br />

will require designing systems where the inherent activities and distribution of nanoparticle size<br />

and composition allow for minimal modifications that will translate into production process steps.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 37<br />

REFERENCES<br />

1. Kayser, O., Lemke, A., and Hernandez-Trejo, N., The impact of nanobiotechnology on the development<br />

of new drug delivery systems, Curr. Pharm. Biotechnol., 6(1), 3–5, 2005.<br />

2. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2(5), 347–360, 2003.<br />

3. Douglas, S. J., Davis, S. S., and Illum, L., Nanoparticles in drug delivery, Crit. Rev. Ther. Drug<br />

Carrier Syst., 3(3), 233–261, 1987.<br />

4. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Nanomedicine: Current status and future<br />

prospects, FASEB J., 19(3), 311–330, 2005.<br />

5. Zimmer, A., Antisense oligonucleotide delivery with polyhexylcyanoacrylate nanoparticles as<br />

carriers, Methods, 18(3), 286–295, 1999, see also p. 322.<br />

6. Li, H. and Qian, Z. M., Transferrin/transferrin receptor-mediated drug delivery, Med. Res. Rev.,<br />

22(3), 225–250, 2002.<br />

7. Balthasar, S. et al., Preparation and characterisation of antibody modified gelatin nanoparticles as<br />

drug carrier system for uptake in lymphocytes, Biomaterials, 26(15), 2723–2732, 2005.<br />

8. Mao, S. J. et al., Uptake of albumin nanoparticle surface modified with glycyrrhizin <strong>by</strong> primary<br />

cultured rat hepatocytes, World J. Gastroenterol., 11(20), 3075–3079, 2005.<br />

9. Kickhoefer, V. A. et al., Engineering of vault nanocapsules with enzymatic and fluorescent properties,<br />

Proc. Natl. Acad. Sci. USA, 102(12), 4348–4352, 2005.<br />

10. Benns, J. M. and Kim, S. W., Tailoring new gene delivery designs for specific targets, J. Drug<br />

Target., 8(1), 1–12, 2000.<br />

11. Prabha, S. and Labhasetwar, V., Nanoparticle-mediated wild-type p53 gene delivery results in<br />

sustained antiproliferative activity in breast cancer cells, Mol. Pharmacol., 1(3), 211–219, 2004.<br />

12. Bala, I., Hariharan, S., and Kumar, M. N., PLGA nanoparticles in drug delivery: The state of the art,<br />

Crit. Rev. Ther. Drug Carrier Syst., 21(5), 387–422, 2004.<br />

13. Bejjani, R. A. et al., Nanoparticles for gene delivery to retinal pigment epithelial cells, Mol. Vis., 11,<br />

124–132, 2005.<br />

14. Olivier, J. C., Drug transport to brain with targeted nanoparticles, NeuroRx, 2(1), 108–119, 2005.<br />

15. Daugherty, A. L. et al., Pharmacological modulation of the tissue response to implanted polylacticco-glycolic<br />

acid microspheres, Eur. J. Pharm. Biopharm., 44(1637), 89–102, 1997.<br />

16. Lemarchand, C. et al., Novel polyester–polysaccharide nanoparticles, Pharm. Res., 20(8),<br />

1284–1292, 2003.<br />

17. Mainardes, R. M. et al., Colloidal carriers for ophthalmic drug delivery, Curr. Drug Targets, 6(3),<br />

363–371, 2005.<br />

18. Kojima, C. et al., Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and<br />

their ability to encapsulate anti-cancer drugs, Bioconjugate Chem., 11(6), 910–917, 2000.<br />

19. Jain, T. K. et al., Iron oxide nanoparticles for sustained delivery of anti-cancer agents, Mol. Pharmacol.,<br />

2(3), 194–205, 2005.<br />

20. He, Q. et al., Calcium phosphate nanoparticle adjuvant, Clin. Diagn. Lab. Immunol., 7(6), 899–903,<br />

2000.<br />

21. Petrovsky, N. and Aguilar, J. C., Vaccine adjuvants: Current state and future trends, Immunol. Cell<br />

Biol., 82(5), 488–496, 2004.<br />

22. Gao, X. et al., In vivo molecular and cellular imaging with quantum dots, Curr. Opin. Biotechnol.,<br />

16(1), 63–72, 2005.<br />

23. Michalet, X. et al., Quantum dots for live cells, in vivo imaging, and diagnostics, Science, 307(5709),<br />

538–544, 2005.<br />

24. Kirchner, C. et al., Cytotoxicity of colloidal CdSe and CdSe/ZnS nanoparticles, Nano Lett., 5(2),<br />

331–338, 2005.<br />

25. Ruysschaert, T. et al., Liposome-based nanocapsules, IEEE Trans. Nanobioscience, 3(1), 49–55,<br />

2004.<br />

26. Shenoy, V. S., Vijay, I. K., and Murthy, R. S., Tumour targeting: Biological factors and formulation<br />

advances in injectable lipid nanoparticles, J. Pharm. Pharmacol., 57(4), 411–422, 2005.<br />

27. Moghimi, S. M. and Hunter, A. C., Capture of stealth nanoparticles <strong>by</strong> the body’s defences, Crit. Rev.<br />

Ther. Drug Carrier Syst., 18(6), 527–550, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


38<br />

Nanotechnology for Cancer Therapy<br />

28. Allen, C. et al., Controlling the physical behavior and biological performance of liposome formulations<br />

through use of surface grafted poly(ethylene glycol), Biosci. Rep., 22(2), 225–250, 2002.<br />

29. Muller, R. H. and Keck, C. M., Challenges and solutions for the delivery of biotech drugs—A review<br />

of drug nanocrystal technology and lipid nanoparticles, J. Biotechnol., 113(1–3), 151–170, 2004.<br />

30. Emanuel, N. et al., Preparation and characterization of doxorubicin-loaded sterically stabilized<br />

immunoliposomes, Pharm. Res., 13(3), 352–359, 1996.<br />

31. Versluis, A. J. et al., Stable incorporation of a lipophilic daunorubicin prodrug into apolipoprotein<br />

E-exposing liposomes induces uptake of prodrug via low-density lipoprotein receptor in vivo,<br />

J. Pharmacol. Exp. Ther., 289(1), 1–7, 1999.<br />

32. Nobs, L. et al., Poly(lactic acid) nanoparticles labeled with biologically active Neutravidin for active<br />

targeting, Eur. J. Pharm. Biopharm., 58(3), 483–490, 2004.<br />

33. Gref, R. et al., Surface-engineered nanoparticles for multiple ligand coupling, Biomaterials, 24(24),<br />

4529–4537, 2003.<br />

34. Konda, S. D. et al., Specific targeting of folate-dendrimer MRI contrast agents to the high affinity<br />

folate receptor expressed in ovarian tumor xenografts, Magma, 12(2–3), 104–113, 2001.<br />

35. Kobayashi, H. and Brechbiel, M. W., Dendrimer-based macromolecular MRI contrast agents:<br />

Characteristics and application, Mol. Imaging, 2(1), 1–10, 2003.<br />

36. McIntyre, J. O. et al., Development of a novel fluorogenic proteolytic beacon for in vivo detection<br />

and imaging of tumour-associated matrix metalloproteinase-7 activity, Biochem. J., 377(Pt 3),<br />

617–628, 2004.<br />

37. Morawski, A. M., Lanza, G. A., and Wickline, S. A., Targeted contrast agents for magnetic resonance<br />

imaging and ultrasound, Curr. Opin. Biotechnol., 16(1), 89–92, 2005.<br />

38. Voura, E. B. et al., Tracking metastatic tumor cell extravasation with quantum dot nanocrystals and<br />

fluorescence emission-scanning microscopy, Nat. Med., 10(9), 993–998, 2004.<br />

39. Marsh, J. N. et al., Improvements in the ultrasonic contrast of targeted perfluorocarbon nanoparticles<br />

using an acoustic transmission line model, IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 49(1),<br />

29–38, 2002.<br />

40. Kiser, P. F., Wilson, G., and Needham, D., A synthetic mimic of the secretory granule for drug<br />

delivery, Nature, 394(6692), 459–462, 1998.<br />

41. Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancer therapy,<br />

Adv. Drug Deliv. Rev., 56(11), 1649–1659, 2004.<br />

42. Mintorovitch, J. and Shamsi, K., Eovist injection and Resovist injection: Two new liver-specific<br />

contrast agents for MRI, Oncology (Williston Park), 14(6 Suppl. 3), 37–40, 2000.<br />

43. Perez, R. V. et al., Selective targeting of Kupffer cells with liposomal butyrate augments portal<br />

venous transfusion-induced immunosuppression, Transplantation, 65(10), 1294–1298, 1998.<br />

44. Kim, B. K. et al., Hydrophilized poly(lactide-co-glycolide) nanospheres with poly(ethylene oxide)–<br />

Poly(propylene oxide)–poly(ethylene oxide) triblock copolymer, J. Microencapsul., 21(7), 697–707,<br />

2004.<br />

45. Veronese, F. M. and Pasut, G., PEGylation, successful approach to drug delivery, Drug Discov.<br />

Today, 10(21), 1451–1458, 2005.<br />

46. Cole, C. et al., Tumor-targeted, systemic delivery of therapeutic viral vectors using hitchhiking on<br />

antigen-specific T cells, Nat. Med., 11(10), 1073–1081, 2005.<br />

47. Maeda, H. et al., Tumor vascular permeability and the EPR effect in macromolecular therapeutics: A<br />

review, J. Control. Release, 65(1–2), 271–284, 2000.<br />

48. Witvrouw, M. et al., Polyanionic (i.e., polysulfonate) dendrimers can inhibit the replication of human<br />

immunodeficiency virus <strong>by</strong> interfering with both virus adsorption and later steps (reverse transcriptase/integrase)<br />

in the virus replicative cycle, Mol. Pharmacol., 58(5), 1100–1108, 2000.<br />

49. Moynagh, P. N., TLR signalling and activation of IRFs: Revisiting old friends from the NF-kappaB<br />

pathway, Trends Immunol., 26(9), 469–476, 2005.<br />

50. Rae, C. S. et al., Systemic trafficking of plant virus nanoparticles in mice via the oral route, Virology,<br />

343(2), 224–235, 2005.<br />

51. Aucouturier, J., Dupuis, L., and Ganne, V., Adjuvants designed for veterinary and human vaccines,<br />

Vaccine, 19(17–19), 2666–2672, 2001.<br />

52. McGuire, M. J. et al., A library-selected, Langerhans cell-targeting peptide enhances an immune<br />

response, DNA Cell Biol., 23(11), 742–752, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 39<br />

53. Mizukami, Y. et al., Induction of interleukin-8 preserves the angiogenic response in HIF-1alphadeficient<br />

colon cancer cells, Nat. Med., 11(9), 992–997, 2005.<br />

54. Donaldson, K. and Tran, C. L., Inflammation caused <strong>by</strong> particles and fibers, Inhal. Toxicol., 14(1),<br />

5–27, 2002.<br />

55. Wottrich, R., Diabate, S., and Krug, H. F., Biological effects of ultrafine model particles in human<br />

macrophages and epithelial cells in mono- and co-culture, Int. J. Hyg. Environ. Health, 207(4),<br />

353–361, 2004.<br />

56. Gopalan, B. et al., Nanoparticle based systemic gene therapy for lung cancer: Molecular mechanisms<br />

and strategies to suppress nanoparticle-mediated inflammatory response, Technol. Cancer Res.<br />

Treat., 3(6), 647–657, 2004.<br />

57. Oberdorster, G., Oberdorster, E., and Oberdorster, J., Nanotoxicology: An emerging discipline<br />

evolving from studies of ultrafine particles, Environ. Health Perspect., 113(7), 823–839, 2005.<br />

58. Segura, S. et al., Potential of albumin nanoparticles as carriers for interferon gamma, Drug Dev. Ind.<br />

Pharm., 31(3), 271–280, 2005.<br />

59. Hamaguchi, T. et al., NK105, a paclitaxel-incorporating micellar nanoparticle formulation, can<br />

extend in vivo antitumour activity and reduce the neurotoxicity of paclitaxel, Br. J. Cancer,<br />

92(7), 1240–1246, 2005.<br />

60. Shenoy, D. B. and Amiji, M. M., Poly(ethylene oxide)-modified poly(epsilon-caprolactone) nanoparticles<br />

for targeted delivery of tamoxifen in breast cancer, Int. J. Pharm., 293(1–2), 261–270,<br />

2005.<br />

61. Park, J. W., Benz, C. C., and Martin, F. J., Future directions of liposome- and immunoliposomebased<br />

cancer therapeutics, Semin. Oncol., 31(6 Suppl. 13), 196–205, 2004.<br />

62. Rose, P. G., Pegylated liposomal doxorubicin: Optimizing the dosing schedule in ovarian cancer,<br />

Oncologist, 10(3), 205–214, 2005.<br />

63. Robert, N. J. et al., The role of the liposomal anthracyclines and other systemic therapies in the<br />

management of advanced breast cancer, Semin. Oncol., 31(6 Suppl. 13), 106–146, 2004.<br />

64. Malik, N., Evagorou, E. G., and Duncan, R., Dendrimer-platinate: A novel approach to cancer<br />

chemotherapy, Anti-cancer Drugs, 10(8), 767–776, 1999.<br />

65. Kobayashi, H. et al., Monoclonal antibody–dendrimer conjugates enable radiolabeling of antibody<br />

with markedly high specific activity with minimal loss of immunoreactivity, Eur. J. Nucl. Med.,<br />

27(9), 1334–1339, 2000.<br />

66. Lam, C. W. et al., Pulmonary toxicity of single-wall carbon nanotubes in mice 7 and 90 days after<br />

intratracheal instillation, Toxicol. Sci., 77(1), 126–134, 2004.<br />

67. Liu, J. J. et al., Multiclass cancer classification and biomarker discovery using GA-based algorithms,<br />

Bioinformatics, 21(11), 2691–2697, 2005.<br />

68. Ewer, M. S. et al., Cardiac safety of liposomal anthracyclines, Semin. Oncol., 31(6 Suppl. 13),<br />

161–181, 2004.<br />

69. Harlos, J. P. and Weiss, L., Differences in the peripheries of Lewis lung tumor cells growing in<br />

different sites in the mouse, Int. J. Cancer, 32(6), 745–750, 1983.<br />

70. Subjeck, J. R., Weiss, L., and Warren, L., Colloidal iron hydroxide-binding to the surfaces of chick<br />

embryo fibroblasts transformed <strong>by</strong> wild-type and a temperature-sensitive mutant of Rous sarcoma<br />

virus, J. Cell. Physiol., 91(3), 329–334, 1977.<br />

71. Lynch, C. C. et al., MMP-7 promotes prostate cancer-induced osteolysis via the solubilization of<br />

RANKL, Cancer Cell, 7(5), 485–496, 2005.<br />

72. Carter, P., Smith, L., and Ryan, M., Identification and validation of cell surface antigens for antibody<br />

targeting in oncology, Endocr. Relat. Cancer, 11(4), 659–687, 2004.<br />

73. Park, J. W. et al., Development of anti-p185HER2 immunoliposomes for cancer therapy, Proc. Natl<br />

Acad. Sci. USA, 92(5), 1327–1331, 1995.<br />

74. Dinauer, N. et al., Selective targeting of antibody-conjugated nanoparticles to leukemic cells and<br />

primary T-lymphocytes, Biomaterials, 26(29), 5898–5906, 2005.<br />

75. Brown, S. L., Miller, R. A., and Levy, R., Antiidiotype antibody therapy of B-cell lymphoma, Semin.<br />

Oncol., 16(3), 199–210, 1989.<br />

76. Cui, Z., Han, S. J., and Huang, L., Coating of mannan on LPD particles containing HPV E7 peptide<br />

significantly enhances immunity against HPV-positive tumor, Pharm. Res., 21(6), 1018–1025, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


40<br />

Nanotechnology for Cancer Therapy<br />

77. Elamanchili, P. et al., Characterization of poly(D,L- lactic-co-glycolic acid) based nanoparticulate<br />

system for enhanced delivery of antigens to dendritic cells, Vaccine, 22(19), 2406–2412, 2004.<br />

78. Mrsny, R. J. et al., Bacterial toxins as tools for mucosal vaccination, Drug Discov. Today, 7(4),<br />

247–258, 2002.<br />

79. Vitols, S., Uptake of low-density lipoprotein <strong>by</strong> malignant cells—Possible therapeutic applications,<br />

Cancer Cells, 3(12), 488–495, 1991.<br />

80. Howarth, M. et al., Targeting quantum dots to surface proteins in living cells with biotin ligase, Proc.<br />

Natl Acad. Sci. USA, 102(21), 7583–7588, 2005.<br />

81. Yu, D. et al., The specific delivery of proteins to human liver cells <strong>by</strong> engineered bio-nanocapsules,<br />

FEBS J., 272(14), 3651–3660, 2005.<br />

82. Popielarski, S. R., Pun, S. H., and Davis, M. E., A nanoparticle-based model delivery system to guide<br />

the rational design of gene delivery to the liver. 1. Synthesis and characterization, Bioconjugate<br />

Chem., 16(5), 1063–1070, 2005.<br />

83. Ma, M. K. and McLeod, H. L., Lessons learned from the irinotecan metabolic pathway, Curr. Med.<br />

Chem., 10(1), 41–49, 2003.<br />

84. Rossin, R. et al., 64 Cu-labeled folate-conjugated shell cross-linked nanoparticles for tumor imaging<br />

and radiotherapy: Synthesis, radiolabeling, and biologic evaluation, J. Nucl. Med., 46(7),<br />

1210–1218, 2005.<br />

85. Kukowska-Latallo, J. F. et al., Nanoparticle targeting of anti-cancer drug improves therapeutic<br />

response in animal model of human epithelial cancer, Cancer Res., 65(12), 5317–5324, 2005.<br />

86. Thomas, T. P. et al., Targeting and inhibition of cell growth <strong>by</strong> an engineered dendritic nanodevice,<br />

J. Med. Chem., 48(11), 3729–3735, 2005.<br />

87. Sonvico, F. et al., Folate-conjugated iron oxide nanoparticles for solid tumor targeting as potential<br />

specific magnetic hyperthermia mediators: Synthesis, physicochemical characterization, and in vitro<br />

experiments, Bioconjugate Chem., 16(5), 1181–1188, 2005.<br />

88. Stevens, P. J., Sekido, M., and Lee, R. J., A folate receptor-targeted lipid nanoparticle formulation<br />

for a lipophilic paclitaxel prodrug, Pharm. Res., 21(12), 2153–2157, 2004.<br />

89. Yang, P. H. et al., Transferrin-mediated gold nanoparticle cellular uptake, Bioconjugate Chem.,<br />

16(3), 494–496, 2005.<br />

90. Bellocq, N. C. et al., Transferrin-containing, cyclodextrin polymer-based particles for tumortargeted<br />

gene delivery, Bioconjugate Chem., 14(6), 1122–1132, 2003.<br />

91. Gonzalez, D. E. et al., An oligopeptide transporter is expressed at high levels in the pancreatic<br />

carcinoma cell lines AsPc-1 and Capan-2, Cancer Res., 58(3), 519–525, 1998.<br />

92. Luo, J. L., Kamata, H., and Karin, M., IKK/NF-kappaB signaling: Balancing life and death—A new<br />

approach to cancer therapy, J. Clin. Invest., 115(10), 2625–2632, 2005.<br />

93. Kelly, K. A. et al., Detection of vascular adhesion molecule-1 expression using a novel multimodal<br />

nanoparticle, Circ. Res., 96(3), 327–336, 2005.<br />

94. Li, K. C., Guccione, S., and Bednarski, M. D., Combined vascular targeted imaging and therapy: A<br />

paradigm for personalized treatment, J. Cell. Biochem. Suppl., 39, 65–71, 2002.<br />

95. Torchilin, V. P., Drug targeting, Eur. J. Pharm. Sci., 11(Suppl. 2), S81–S91, 2000.<br />

96. Gabizon, A. A., Selective tumor localization and improved therapeutic index of anthracyclines<br />

encapsulated in long-circulating liposomes, Cancer Res., 52(4), 891–896, 1992.<br />

97. Allen, T. M., Long-circulating (sterically stabilized) liposomes for targeted drug delivery, Trends<br />

Pharmacol. Sci., 15(7), 215–220, 1992.<br />

98. Backer, M. V. et al., Vascular endothelial growth factor selectively targets boronated dendrimers to<br />

tumor vasculature, Mol. Cancer Ther., 4(9), 1423–1429, 2005.<br />

99. Li, L. et al., A novel antiangiogenesis therapy using an integrin antagonist or anti-Flk-1 antibody<br />

coated 90 Y-labeled nanoparticles, Int. J. Radiat. Oncol. Biol. Phys., 58(4), 1215–1227, 2004.<br />

100. Konstan, M. W. et al., Compacted DNA nanoparticles administered to the nasal mucosa of cystic<br />

fibrosis subjects are safe and demonstrate partial to complete cystic fibrosis transmembrane regulator<br />

reconstitution, Hum. Gene Ther., 15(12), 1255–1269, 2004.<br />

101. Sengupta, S. et al., Temporal targeting of tumour cells and neovasculature with a nanoscale delivery<br />

system, Nature, 436(7050), 568–572, 2005.<br />

102. Tanaka, K. et al., Intratumoral injection of immature dendritic cells enhances antitumor effect of<br />

hyperthermia using magnetic nanoparticles, Int. J. Cancer, 116(4), 624–633, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Active Targeting Strategies in Cancer with a Focus on Potential Nanotechnology Applications 41<br />

103. Tsourkas, A. et al., In vivo imaging of activated endothelium using an anti-VCAM-1 magnetooptical<br />

probe, Bioconjugate Chem., 16(3), 576–581, 2005.<br />

104. Pasqualini, R. and Ruoslahti, E., Organ targeting in vivo using phage display peptide libraries,<br />

Nature, 380(6572), 364–366, 1996.<br />

105. Akerman, M. E. et al., Nanocrystal targeting in vivo, Proc. Natl Acad. Sci. USA, 99(20),<br />

12617–12621, 2002.<br />

106. Bib<strong>by</strong>, D. C. et al., Pharmacokinetics and biodistribution of RGD-targeted doxorubicin-loaded<br />

nanoparticles in tumor-bearing mice, Int. J. Pharm., 293(1–2), 281–290, 2005.<br />

107. Hood, J. D. et al., Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science,<br />

296(5577), 2404–2407, 2002.<br />

108. Lamprecht, A. et al., Biodegradable nanoparticles for targeted drug delivery in treatment of inflammatory<br />

bowel disease, J. Pharmacol. Exp. Ther., 299(2), 775–781, 2001.<br />

109. Lamprecht, A. et al., A pH-sensitive microsphere system for the colon delivery of tacrolimus<br />

containing nanoparticles, J. Control. Release, 104(2), 337–346, 2005.<br />

110. Hussain, M. et al., A novel anionic dendrimer for improved cellular delivery of antisense oligonucleotides,<br />

J. Control. Release, 99(1), 139–155, 2004.<br />

111. Santos Maia, C. et al., Drug targeting <strong>by</strong> solid lipid nanoparticles for dermal use, J. Drug Target.,<br />

10(6), 489–495, 2002.<br />

112. Alvarez-Roman, R. et al., Skin penetration and distribution of polymeric nanoparticles, J. Control.<br />

Release, 99(1), 53–62, 2004.<br />

113. Prausnitz, M. R. et al., Microneedles for transdermal drug delivery, Adv. Drug Deliv. Rev., 56(5),<br />

581–587, 2004.<br />

114. Bourges, J. L. et al., Ocular drug delivery targeting the retina and retinal pigment epithelium using<br />

polylactide nanoparticles, Invest. Ophthalmol. Vis. Sci., 44(8), 3562–3569, 2003.<br />

115. Florence, A. T., Issues in oral nanoparticle drug carrier uptake and targeting, J. Drug Target., 12(2),<br />

65–70, 2004.<br />

116. Florence, A. T. et al., Factors affecting the oral uptake and translocation of polystyrene nanoparticles:<br />

Histological and analytical evidence, J. Drug Target., 3(1), 65–70, 1995.<br />

117. Takeuchi, H. et al., Effectiveness of submicron-sized, chitosan-coated liposomes in oral administration<br />

of peptide drugs, Int. J. Pharm., 303(1–2), 160–170, 2005.<br />

118. Hu, J. et al., A remarkable permeability of canalicular tight junctions might facilitate retrograde,<br />

non-viral gene delivery to the liver via the bile duct, Gut, 54(10), 1473–1479, 2005.<br />

119. Yoncheva, K., Lizarraga, E., and Irache, J. M., Pegylated nanoparticles based on poly(methyl vinyl<br />

ether-co-maleic anhydride): Preparation and evaluation of their bioadhesive properties, Eur.<br />

J. Pharm. Sci., 24(5), 411–419, 2005.<br />

120. Wiwattanapatapee, R. et al., Anionic PAMAM dendrimers rapidly cross adult rat intestine in vitro: A<br />

potential oral delivery system?, Pharm. Res., 17(8), 991–998, 2000.<br />

121. Jevprasesphant, R. et al., Engineering of dendrimer surfaces to enhance transepithelial transport and<br />

reduce cytotoxicity, Pharm. Res., 20(10), 1543–1550, 2003.<br />

122. Sham, J. O. et al., Formulation and characterization of spray-dried powders containing nanoparticles<br />

for aerosol delivery to the lung, Int. J. Pharm., 269(2), 457–467, 2004.<br />

123. Borm, P. J. and Kreyling, W., Toxicological hazards of inhaled nanoparticles—Potential implications<br />

for drug delivery, J. Nanosci. Nanotechnol., 4(5), 521–531, 2004.<br />

124. Liu, G. et al., Nanoparticle and other metal chelation therapeutics in Alzheimer disease, Biochim.<br />

Biophys. Acta, 1741(3), 246–252, 2005.<br />

125. Muller, R. H. and Keck, C. M., Drug delivery to the brain—Realization <strong>by</strong> novel drug carriers,<br />

J. Nanosci. Nanotechnol., 4(5), 471–483, 2004.<br />

126. Garcia-Garcia, E. et al., A relevant in vitro rat model for the evaluation of blood–brain barrier<br />

translocation of nanoparticles, Cell. Mol. Life Sci., 62(12), 1400–1408, 2005.<br />

127. Sun, W. et al., Specific role of polysorbate 80 coating on the targeting of nanoparticles to the brain,<br />

Biomaterials, 25(15), 3065–3071, 2004.<br />

128. Thorne, R. G. et al., Quantitative analysis of the olfactory pathway for drug delivery to the brain,<br />

Brain Res., 692(1–2), 278–282, 1995.<br />

129. Vila, A. et al., Transport of PLA–PEG particles across the nasal mucosa: Effect of particle size and<br />

PEG coating density, J. Control. Release, 98(2), 231–244, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


42<br />

Nanotechnology for Cancer Therapy<br />

130. Moghimi, S. M. and Szebeni, J., Stealth liposomes and long circulating nanoparticles: Critical issues<br />

in pharmacokinetics, opsonization and protein-binding properties, Prog. Lipid Res., 42(6), 463–478,<br />

2003.<br />

131. Ma, Z. and Lim, L. Y., Uptake of chitosan and associated insulin in Caco-2 cell monolayers:<br />

A comparison between chitosan molecules and chitosan nanoparticles, Pharm. Res., 20(11),<br />

1812–1819, 2003.<br />

132. Gumbleton, M. et al., Targeting caveolae for vesicular drug transport, J. Control. Release, 87(1–3),<br />

139–151, 2003.<br />

133. Qaddoumi, M. G. et al., Clathrin and caveolin-1 expression in primary pigmented rabbit conjunctival<br />

epithelial cells: Role in PLGA nanoparticle endocytosis, Mol. Vis., 9, 559–568, 2003.<br />

134. Dass, C. R., Oligonucleotide delivery to tumours using macromolecular carriers, Biotechnol. Appl.<br />

Biochem., 40(Pt 2), 113–122, 2004.<br />

135. Gupta, A. K. and Curtis, A. S., Lactoferrin and ceruloplasmin derivatized superparamagnetic iron<br />

oxide nanoparticles for targeting cell surface receptors, Biomaterials, 25(15), 3029–3040, 2004.<br />

136. de la Fuente, J. M. and Berry, C. C., Tat Peptide as an efficient molecule to translocate gold<br />

nanoparticles into the cell nucleus, Bioconjugate Chem., 16(5), 1176–1180, 2005.<br />

137. Li, N. et al., Ultrafine particulate pollutants induceoxidative stress and mitochondrial damage,<br />

Environ. Health Perspect., 111(4), 455–460, 2003.<br />

138. Savic, R. et al., Micellar nanocontainers distribute to defined cytoplasmic organelles, Science,<br />

300(5619), 615–618, 2003.<br />

139. Fischer, H. W., Improvement in radiographic contrast media through the development of colloidal or<br />

particulate media: An analysis, J. Theor. Biol., 67(4), 653–670, 1977.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


4<br />

CONTENTS<br />

Pharmacokinetics of<br />

Nanocarrier-Mediated Drug<br />

and Gene Delivery<br />

Yuriko Higuchi, Shigeru Kawakami, and Mitsuru Hashida<br />

4.1 Introduction ...........................................................................................................................44<br />

4.2 Characteristics of the Structure and Pharmacokinetic Properties<br />

of the Nanocarrier .................................................................................................................44<br />

4.2.1 Linear Polymers ........................................................................................................44<br />

4.2.1.1 Their History and Characteristics ..............................................................44<br />

4.2.1.2 Pharmacokinetic Properties and Applications of Drug Carriers ...............45<br />

4.2.2 Polymeric Micelles....................................................................................................46<br />

4.2.2.1 Their History and Characteristics ..............................................................46<br />

4.2.2.2 Pharmacokinetic Properties and Applications as Drug Carriers ...............46<br />

4.2.3 Dendrimers ................................................................................................................47<br />

4.2.3.1 Their History and Characteristics ..............................................................47<br />

4.2.3.2 Pharmacokinetic Properties and Applications as Drug Carriers ...............47<br />

4.2.4 Liposomes..................................................................................................................48<br />

4.2.4.1 Their History and Characteristics ..............................................................48<br />

4.2.4.2 Pharmacokinetic Properties and Applications of Drug Carriers ...............49<br />

4.3 Pharmacokinetic Analysis and Tissue Distribution Characteristics<br />

of Macromolecules ................................................................................................................50<br />

4.3.1 Effect of Size on the Pharmacokinetic Properties ....................................................51<br />

4.3.2 Effect of Charge on the Pharmacokinetic Properties ...............................................51<br />

4.3.3 Effect of Modification for Cell-Selective Delivery ..................................................51<br />

4.4 Nanocarriers for Gene Delivery............................................................................................51<br />

4.4.1 Polymers ....................................................................................................................52<br />

4.4.2 Liposomes..................................................................................................................52<br />

4.4.2.1 Cationic Liposomes for Lung Targeting ...................................................52<br />

4.4.2.2 Mannose Liposomes for Liver NPC Targeting .........................................53<br />

4.4.2.3 Galactose Liposomes for Liver PC Targeting...........................................53<br />

4.4.2.4 Folate Receptor-Mediated Targeting .........................................................54<br />

References ......................................................................................................................................54<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

43


44<br />

4.1 INTRODUCTION<br />

The recent development of nanoscale technologies is beginning to change the foundations of<br />

disease prevention, diagnosis, and treatment. Nanotechnology has allowed improvement of a<br />

novel drug carrier, nanocarrier, involving the nanoscale size and capable of targeting different<br />

cells in the body. These include polymeric micelles, dendrimers, and liposomes. The focus is on<br />

anti-tumor drugs and, although various anti-tumor drugs have been developed for cancer<br />

chemotherapy, such drugs often cause severe side effects related to the high cytotoxicity to<br />

tumor cells, and they are often toxic to normal cells. Therefore, a carefully designed nanocarrier<br />

is required to deliver anti-tumor drugs to their target sites for effective chemotherapy.<br />

Tissue accumulation and cellular uptake of externally administered agents are generally<br />

determined <strong>by</strong> their physicochemical and biological properties and the anatomical and physiological<br />

properties of the body or tissues. As far as cancer therapy is concerned, site-specific delivery of<br />

drugs <strong>by</strong> carrier is broadly categorized as passive and active targeting. Passive targeting is the<br />

method determined <strong>by</strong> the physicochemical properties of the carrier relative to the anatomical and<br />

physiological characteristics of the tissue. From the 1970s forward, it has been demonstrated that<br />

macromolecule–drug conjugates accumulate in high concentrations in solid tumors and produce a<br />

therapeutic effect on the cancer. 1–4 This phenomenon is dependent on the anatomical and physiological<br />

characteristics of the solid tumor tissue, including a large vascular permeability, high<br />

interstitial diffusivity, and a lack of lymphatic drainage. 5,6 As far as this effect is concerned,<br />

Matsumura and Maeda proposed the concept of an enhanced permeability and retention<br />

effect (EPR effect) of macromolecules in solid tumors. 7,8 On the other hand, active targeting<br />

refers to alterations in the drug carrier, using specific interactions such as ligand–receptor and<br />

antigen–antibody phenomena. This approach involves designing carrier materials such as lipids<br />

and polymers and the combination of materials that enable site-specific drug delivery at a cellular or<br />

sub-cellular level.<br />

Because the therapeutic effect of anti-tumor drugs is closely related to their pharmacokinetic<br />

properties, it is important for the rational design of a nanocarrier for targeted delivery to understand<br />

the pharmacokinetics of the nanocarrier in relation to the physicochemical and biological properties<br />

of the drugs involved. 9,10 In this chapter, site-specific delivery using nanocarriers is discussed based<br />

on pharmacokinetic considerations. As far as tissue distribution is concerned, because a nanocarrier<br />

possesses several features common to the macromolecule, the pharmacokinetics properties of the<br />

macromolecule are also summarized.<br />

4.2 CHARACTERISTICS OF THE STRUCTURE AND PHARMACOKINETIC<br />

PROPERTIES OF THE NANOCARRIER<br />

Recently, progress in nanotechnology has allowed the development of several types of nanocarriers<br />

capable of delivering drugs to target tissue and cells. Figure 4.1 summarizes typical types of<br />

nanocarriers and their characteristic structures. Because each type of nanocarrier has a unique<br />

structure and physicochemical characteristics, it exhibits unique pharmacokinetic properties in<br />

the body. In order to develop a strategy for establishing a nanocarrier system, it is necessary to<br />

understand these pharmacokinetic properties.<br />

4.2.1 LINEAR POLYMERS<br />

4.2.1.1 Their History and Characteristics<br />

Nanotechnology for Cancer Therapy<br />

In recent decades, the use of polymers as carriers of both covalently bound and physically entrapped<br />

drugs has been widely explored. The larger hydrodynamic volume of polymers contributes to the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 45<br />

Polymer conjugates<br />

Drug<br />

Dendrimers<br />

Polymer<br />

backbone<br />

(a) (b)<br />

(PAMAM dendrimers)<br />

N<br />

NH2 NH2 (c) (d)<br />

FIGURE 4.1 Typical nanocarriers.<br />

Ethylenediamine<br />

Methyl acrylate<br />

Polymeric micelles<br />

Liposomes<br />

increase in the plasma half-life of drug–polymer conjugates, increasing the probability of accumulation<br />

of the therapeutic agent in the tumor tissue because of the EPR effect. Drug–polymer<br />

conjugates (Figure 4.1) exhibit improved water solubility and reduced toxicity as a result of<br />

accumulation in the target tissue, and they protect the drug from enzymatic degradation or<br />

hydrolysis. However, drugs covalently bound to a polymer lose their pharmacological activity<br />

because a drug cannot interact with the target site because of steric constraints. In this respect,<br />

in 1975, Ringesdorf 11 was the first to propose a model for the rational design of a polymer conjugate<br />

with a drug that consisted of five components: the polymeric backbone, the drug, the spacer, the<br />

targeting group, and the solubilizing agent. The polymeric carrier can be either an inert or a<br />

biodegradable polymer. The drug can be fixed directly or via a spacer group onto the polymer<br />

backbone that is about 5–40 kDa. The proper selection of this spacer offers the possibility of<br />

controlling the site and the rate of release of the active drug from the conjugate <strong>by</strong> hydrolytic or<br />

enzymatic cleavage.<br />

4.2.1.2 Pharmacokinetic Properties and Applications of Drug Carriers<br />

Hydrophobic<br />

polymer<br />

Hydrophilic<br />

polymer<br />

Phospholipid<br />

Cholesterol<br />

To date, poly[N-(2-hydroxypropyl)methacrylamide] (PHPMA) is the most advanced example of a<br />

synthetic polymer used as a drug carrier in both basic research and clinical applications reported <strong>by</strong><br />

Duncan 12 and Kocheck 13 . The biocompatibility of this polymer has been well characterized, 14 and<br />

it was found to have an inert character when it was injected in the blood stream. 15,16 A PHPMA<br />

copolymer conjugated with doxorubicin as an anti-tumor agent linked via a peptidyl linker<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


46<br />

(Gly–Phe–Leu–Gly) was developed. The PHPMA copolymer conjugated with doxorubicin is<br />

stable in plasma 17 and has been shown to be concentrated within solid tumor models. 18,19 It is<br />

then cleaved intracellulary <strong>by</strong> lysosomal cysteine proteinases, 20 there<strong>by</strong> allowing intratumoral<br />

drug release. Preclinical investigations have shown that a PHPMA copolymer conjugated with<br />

doxorubicin has radically different pharmacokinetics compared to free doxorubicin with a distribution<br />

plasma half-life that is increased from 5 min to 1 h. 21 The stable peptidyl linker also ensures<br />

that little or no free doxorubicin is released into the circulation following intravenous injection,<br />

increasing the therapeutic index of the conjugate. Besides the prolonged circulation in blood and<br />

the high solid tumor accumulation, it is also possible to attach targeting moieties such as<br />

antibody 22 and saccharides 23 to PHPMA. Another polymer conjugate, poly(ethylene glycol)<br />

(PEG), is the most basic and widely used polymer. Stylene maleic anhydride conjugated with<br />

the anti-tumor protein, neocarzinostatin (SMANCS), has been studied and is already used for the<br />

treatment of hepatocellular carcinoma. 24 Furthermore, <strong>by</strong> attaching homing devices as a cellspecific<br />

uptake enhancer, the biodistribution could be altered and cellular targeting of drugs<br />

conjugated with polymer could be achieved.<br />

4.2.2 POLYMERIC MICELLES<br />

4.2.2.1 Their History and Characteristics<br />

Polymeric micelles that are prepared from amphiphilic block or graft copolymers with a spherical<br />

core and a shell with a carrier size of 10–100 nm have been undergoing investigation since 1984<br />

and have been studied more actively since 1990. 25–27 The term block refers to the linear architecture<br />

of the copolymer in which the end of one segment is covalently joined to the head of the other<br />

segment to give a diblock AB type (Figure 4.1) or multiple block (AB n) type copolymers. On the<br />

other hand, graft copolymers have a comb-like structure with hydrophilic segments attached on the<br />

side of the cationic segments. Although the influence of the copolymer architecture on biological<br />

activity has not yet been clarified, both block and graft copolymers can form polymeric micelles<br />

because of their amphiphilic character. Interactions between the polymer chains that serve as the<br />

driving force for micelle formation include hydrophobic, electrostatic, and p–p interactions and<br />

hydrogen bonding. Because hydrophobic drugs can be stably trapped in the hydrophobic core of the<br />

polymeric micelles and exhibit water-solubility, polymeric micelles are attractive carriers for<br />

hydrophobic drugs. Moreover, electrostatic (ionic) interactions involving the hydrophilic surface<br />

of polymeric micelles may also be applicable to macromolecules possessing many electrostatic<br />

(ionic) charges, e.g., DNA. 28<br />

4.2.2.2 Pharmacokinetic Properties and Applications as Drug Carriers<br />

Nanotechnology for Cancer Therapy<br />

Early reports showed that the anti-cancer drug, doxorubicin, could be incorporated into a PEGpolyaspartate<br />

block copolymer. 29,30 PEG constituted the outer shell of the micelle that conferred a<br />

stealth property on the drug. After intravenous injection, doxorubicin incorporated in polymeric<br />

micelles was less avidly taken up <strong>by</strong> the reticuloendothelial system (RES) and could be retained in<br />

the circulation for a longer time compared with an intravenous injection of doxorubicin alone. 29<br />

The in vivo anti-tumor activity against Colon-26 solid tumors showed that there was critical<br />

suppression of tumor growth and a prolonged life span for doxorubicin incorporated in polymeric<br />

micelles when administered to mice. 30<br />

Another type of polymeric micelle composed poly(ethylene oxide–aspartate)block copolymer<br />

and doxorubicin conjugate also exhibited a sustained circulation in blood, 31 reduced uptake <strong>by</strong> the<br />

RES, and more than 100 times higher accumulation in tumors in Colon-26 tumor-bearing mice. 32<br />

Another report described the tumor-targeting delivery of paclitaxel <strong>by</strong> block copolymers producing<br />

both anti-tumor activity and a reduction in a major side effect of paclitaxel, neurotoxicity. 33<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 47<br />

In a previous study, all-trans retinoic acid (ATRA) incorporated polymeric micelles produced a<br />

higher AUC and lower hepatic clearance than free ATRA, demonstrating escape from the RES. 34<br />

Polymeric micelles have also been used in an active targeting approach, using the fact that<br />

certain pathological processes are associated with local temperature increase and/or acidosis. The<br />

use of polymeric micelles prepared from thermo- 35,36 or pH-sensitive 37 block co-polymers allows<br />

the disintegration of micelles and release of the incorporated drugs specifically at the site of interest.<br />

Additionally, the micelles can be targeted <strong>by</strong> attaching to their surface a vector molecule such as an<br />

antibody, peptide, lectin, saccharide, hormone, or some low-molecular weight compounds if this<br />

vector molecule binds to ligands characteristic of the site of interest.<br />

4.2.3 DENDRIMERS<br />

4.2.3.1 Their History and Characteristics<br />

Dendrimers are a new class of polymeric materials discovered in the 1980s that are highly<br />

branched, monodisperse macromolecules. These hyperbranched molecules were described <strong>by</strong><br />

Tomalia et al. 38 and named dendrimers. The word dendrimer comes from the Greek dendron,<br />

meaning a tree or a branch, and meros meaning part. At the same time, Vogtle and his co-workers 39<br />

published the first report of the synthesis of polymers with a dendritic structure that they called<br />

cascade molecules. Newkome et al. 40 independently reported the synthesis of similar macromolecules<br />

that they called arboros from the Latin word arbor, meaning a tree.<br />

Although linear polymers are highly polydisperse with a wide range of molecular weights<br />

because of their automatic assembly, dendrimers are mono dispersed or low-polydispersed that<br />

is necessary for homogeneous and reproducible pharmacokinetic properties. Therefore, dendrimers<br />

that have well-defined structures and flexible functions are good candidates for drug carriers.<br />

Poly(amidoamine) (PAMAM) dendrimers (Figure 4.1) are the first complete dendrimer family<br />

to be synthesized, characterized, and commercialized. Dendrimers have a molecular architecture<br />

characterized <strong>by</strong> regular, dendric branching with radial symmetry. The molecular weights are<br />

similar to those of proteins, and their size ranges from 1 to 10 nm with each generation, or<br />

layer, of polymer adding 1 nm to the diameter of the molecule. 41 Lower generation dendrimers<br />

have a highly asymmetric shape and possess more open structures compared with those of a higher<br />

generation. As the chains growing from the core molecule become longer and more branched,<br />

dendrimers adopt a globular structure. Dendrimers become densely packed as they extend out to the<br />

periphery that results in a closed membrane-like structure. When a critical branched state is<br />

reached, dendrimers cannot continue to grow because of a lack of space. This is called the starburst<br />

effect. In the case of PAMAM dendrimer synthesis, this was observed after the 10th generation that<br />

contains 6141 monomer units and has a diameter of about 12.4 nm. 41<br />

4.2.3.2 Pharmacokinetic Properties and Applications as Drug Carriers<br />

As far as dendrimers’ pharmacokinetics are concerned, the biodistribution of [ 14 C] labeled<br />

PAMAM dendrimers after intravenous injection was determined for generation 3, 5,<br />

and 7 PAMAM dendrimers. 42 At each generation, the biodistribution characteristics were different.<br />

Generation 3 dendrimers showed the highest accumulation in kidney (15% of dose/g tissue over<br />

48 h); however, generation 5 and 7 dendrimers preferentially accumulated in the pancreas<br />

(peak level 32% dose/g tissue at 24 h and 20% of dose/g tissue at 2 h, respectively). In addition,<br />

generation 7 dendrimers showed extremely high urinary excretion with values of 46 and 74% of<br />

dose/g tissue at 2 and 4 h, respectively. These results suggested that the biodistribution of a<br />

dendrimer depended on the generation because physicochemical properties were determined <strong>by</strong><br />

the generation. Other reports described the biodistribution of [ 3 H]-labeled neutral and positively<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


48<br />

charged generation 5 PAMAM dendrimers in a tumor-bearing mouse model. 43 Both neutral and<br />

positively charged dendrimers showed a similar biodistribution trend 1 h after intravenous injection;<br />

the highest level was found in the lungs, kidney, and liver (positive, 28–6.1% of dose/g<br />

tissue; neutral, 5.3–2.9% of dose/g tissue) followed <strong>by</strong> the tumor, spleen, and pancreas (positive,<br />

3.3–2.6% of dose/g, tissue; neutral 3.4–0.92% of dose/g tissue). Differences between neutral and<br />

positively charged dendrimers were observed for accumulation in the lung. A high lung accumulation<br />

of cationic dendrimer was found compared with conventional cationic liposomes, and this<br />

may be explained <strong>by</strong> the stacking to the capillary vessels of the lung. Moreover, [ 14 C]-labeling<br />

of the dendrimer results in partial neutralization of the dendrimer. Therefore, different biodistribution<br />

properties were observed in the lung and liver accumulation between [ 14 C]-and [ 3 H]labeled<br />

dendrimers. These results suggest that the biodistribution properties of radio-labeled<br />

dendrimers have to be examined carefully, taking into consideration the surface charge<br />

caused <strong>by</strong> the radio-labeling. As far as tumor accumulation is concerned, this may be explained<br />

<strong>by</strong> the EPR effect.<br />

Drug delivery applications of dendrimers have been studied <strong>by</strong> Szoka and co-workers. 44 The<br />

biodistribution of doxorubicin covalently bound via a hydrazone linkage with a polyester dendritic<br />

scaffold based on 2,2-bis(hydroxymethyl)propanoic acid was examined. After intravenous injection,<br />

doxorubicin was extracted from the tumor and organs. Compared with free doxorubicin, the<br />

serum half-life was significantly increased. Therefore, this dendrimer would be a promising drug<br />

carrier for cancer therapy.<br />

4.2.4 LIPOSOMES<br />

4.2.4.1 Their History and Characteristics<br />

Nanotechnology for Cancer Therapy<br />

Liposomes that are vesicles of lipid bilayers were first prepared in the 1960s. Liposome encapusulation<br />

has been proven useful for reducing the toxicity of drugs and increasing the solubility of<br />

hydrophobic drugs. However, liposomes tend to be trapped <strong>by</strong> the RES after intravenous injection.<br />

Therefore, allowing escape from the RES is one of the key strategies for developing cell-specific<br />

carriers because liposomes escaping from the RES accumulate in tumor tissue <strong>by</strong> a passive<br />

mechanism. As far as methods of escaping from the RES are concerned, Allen et al. in 1987 45<br />

were the first to show that ganglioside and sphingomyelin (SM) synergistically acted to dramatically<br />

reduce the rate and extent of uptake of liposomes <strong>by</strong> the RES; therefore, this type of liposome<br />

is called a stealth liposome. In 1990, Klibanov et al. 46 demonstrated that PEG-liposomes prepared<br />

as poly (ethylene glycol)–phosphatidylethanolamine conjugates (PEG–PE) could also reduce<br />

uptake <strong>by</strong> the RES. Because PEG is easy to prepare and relatively cheap, and it has controllable<br />

molecular weight and is able to bind to lipids or proteins, it can be used for active targeting such as<br />

glycosylated liposomes 47 and immuno-liposomes. 48<br />

On the other hand, as far as active targeting is concerned, the receptor-mediated endocytosis<br />

systems present in various cell types would be useful, and a number of gene delivery systems were<br />

developed in the 1990s. Ligand–receptor recognition is an attractive tool for the development of<br />

cell-specific targeting systems; i.e., galactose to asialoglycoprotein receptors expressed on hepatocytes<br />

parenchymal cell (PC), 49 mannose to mannose receptors expressed on macrophages,<br />

sinusoidal endothelial cells, Kupffer cells, and dendritic cells, 50 folate receptors expressed on<br />

cancer cells, 51 and so on. As far as the size effect is concerned, it is possible to prepare liposomes<br />

of the size required <strong>by</strong> filtration. Because capillary vessels in a human tumor inoculated into SCID<br />

mice are permeable even to liposomes up to 400 nm in diameter, 52 liposomes with a diameter less<br />

than 50–200 nm are typically used. On the other hand, with regard to the effect of the surface charge<br />

of liposomes, it is known that strongly cationic liposomes highly accumulate in the lung after<br />

intravenous injection. Table 4.1 summarizes liposomes for drug delivery.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 49<br />

TABLE 4.1<br />

In Vivo Targeted Drug Delivery<br />

System Method Target Tissue (Cell) Reference<br />

EPR (prolongation of circulating) Sphingomyelin modification Tumor 45<br />

53<br />

55<br />

56<br />

PEG modification Tumor 46<br />

57<br />

58<br />

59<br />

Asialoglycoprotein receptor mediated Galactose modification Liver (hepatocyte) 62<br />

63<br />

64<br />

61<br />

65<br />

Mannose receptor mediated Mannose modification Liver (non-parenchymal cell) 66<br />

67<br />

71<br />

4.2.4.2 Pharmacokinetic Properties and Applications of Drug Carriers<br />

SM-liposomes described <strong>by</strong> Allen et al. dramatically reduced the uptake <strong>by</strong> the RES and exhibited a<br />

prolonged circulation in blood and, consequently, accumulated in tumor tissue. 45 Gabizon and<br />

Papahadjopoulos 53 have examined a number of other natural and synthetic lipids to prolong the<br />

circulation time and lead to preferential uptake <strong>by</strong> tumor cells in vivo. In these respects, SM-containing<br />

liposomes would be an attractive option as a tumor targeting carrier for highly lipophilic<br />

anti-tumor drugs. The preparation of SM containing emulsion formulations has been also described<br />

to prolong the blood retention of encapsulated [ 3 H] labeled ATRA. 54 Furthermore, after the intravenous<br />

injection of vincristine encapsulated in SM-liposomes, the plasma vincristine level was<br />

7-fold higher than that following administration in the form of bare liposomes, and the half-life of<br />

the elimination from the circulation was prolonged. 55,56 The improved circulation lifetime of<br />

vincristine in SM-liposomes correlated with the increased vincristine accumulation in subcutaneous<br />

solid A431 human xenograft tumors. Moreover, treatment with vincristine in SM-liposomes<br />

delayed the increase in tumor mass. 55<br />

PEG-liposomes are the most widely studied stealth liposomes. This occurs because the presence<br />

of PEG protects liposomes from the interaction with opsonins in the blood plasma and prevents their<br />

rapid uptake <strong>by</strong> the RES. 46,57,58 Therefore, doxorubicin encapsulated into PEG liposomes exhibited<br />

a dramatically increase in circulating time in blood and much better therapeutic efficacy as compared<br />

with the free drug. Therefore, they are currently being used in the clinic as a component of Doxil w . 59<br />

On the other hand, as far as active targeting is concerned, galactose modification could be an<br />

attractive tool for PC targeting of lipophilic drugs. Recently, a novel galactosylated cholesterol<br />

derivative, Gal-C4-Chol, that could be stably incorporated into liposomes <strong>by</strong> means of its lipophilic<br />

anchor and allowed the introduction of galactose residues on the surface of the liposomes was<br />

synthesized. 60,61 After intravenous injection, galactosylated liposomes prepared using Gal-C4-Chol<br />

were preferentially taken up <strong>by</strong> the liver. 62 In addition, prostaglandin E 63 and probucol 64 were<br />

investigated as model lipophilic drugs incorporated in galactosylated liposomes prepared using<br />

Gal-C4-Chol, and they were found to be effectively taken up <strong>by</strong> PC. Recently, other types of<br />

galactosylated liposome were developed. 65<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


50<br />

Using a mannosylated cholesterol derivative (Man-C4-Chol) that was synthesized <strong>by</strong> the same<br />

method as Gal-C4-Chol, mannosylated liposomes (Man-liposomes) for non-parenchymal cell<br />

(NPC) and macrophages targeting were prepared. 66–68 As far as cancer immunotherapy was<br />

concerned, macrophages, Kupffer cells, and dendric cells are attractive targets. These cells<br />

expressed a large number of mannose receptors on the cell surface and, therefore, Man-liposomes<br />

would be ideal carriers for cancer therapy. Previous studies have shown that Kupffer cells<br />

(contained in NPC) can be activated to a tumor reducible state <strong>by</strong> the administration of immunomodulators<br />

such as muramyl dipeptide. 69 However, muramyl dipeptide parentally administrated in<br />

free form is rapidly cleared from the body and excreted in the urine. 70 It was demonstrated that<br />

active targeting of muramyl dipeptide to liver non-parenchymal cells <strong>by</strong> Man-liposomes resulted in<br />

more effective inhibition of liver metastasis than delivery of muramyl dipeptide <strong>by</strong> liposomes<br />

without mannose. 71 Moreover, treatment with muramyl dipeptide incorporated in Man-liposomes<br />

increased the survival of tumor-bearing mice. 71<br />

4.3 PHARMACOKINETIC ANALYSIS AND TISSUE DISTRIBUTION<br />

CHARACTERISTICS OF MACROMOLECULES<br />

The pharmacokinetic characteristics are generally determined <strong>by</strong> the physicochemical properties of<br />

the molecule such as the size, molecular weight, and surface charge. 3 Moreover, specific<br />

interactions such as ligand–receptor or antibody–antigen have a major effect on the characteristics<br />

of the molecule after intravenous injection. 4 The former factors are closely related to passive<br />

targeting, and the latter factors are closely related to active targeting. Figure 4.2 summarizes the<br />

Apparent hepatic uptake clearance (ml/hr)<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

pDNA<br />

PEG-CAT<br />

BSA<br />

(67,000)<br />

PEG-SOD<br />

Hepatic plasma flow rate<br />

Cation:<br />

Electrostatic interaction<br />

Cat-BSA<br />

Weak anion:<br />

Long circulation<br />

Man-SOD<br />

Suc-BSA<br />

Strong anion:<br />

Electrostatic interaction<br />

SOD<br />

(32,000)<br />

0.01 0.1 1<br />

10<br />

Urinary excretion clearance (ml/h)<br />

Gal-SOD<br />

ODN<br />

Inurin<br />

(5,200)<br />

Rate of fluid-phase endocytosis of liver<br />

Nanotechnology for Cancer Therapy<br />

Glomerular filtration rate<br />

Small macromolecule:<br />

Urinary excretion<br />

FIGURE 4.2 Effect of physicochemical characteristics of macromolecules on their hepatic uptake and urinary<br />

excretion in mice. The number in parenthesis is the molecular weight. Abbreviations in this figure are as<br />

follows: SOD, superoxide dismutase; Man-SOD, mannosylated superoxide dismutase; Gal-SOD, galactosylated<br />

superoxide dismutase; PEG-SOD, pegylated superoxide dismutase; Cat-BSA, cationized bovine serum<br />

albumin; PEG-CAT, pegylated catalase; ODN, oligonucleotide. (From Kawabata, K., Takakura, Y., and<br />

Hashida, M., Pharm. Res., 12, 1995; Hashida, M., et al., J. Control. Release, 41,1996; Fujita, T., et al., J.<br />

Pharmacol. Exp. Theor., 263, 1992; Yamasaki, Y., et al., J. Pharmacol. Exp. Theor., 301, 2002; Yabe, Y.,<br />

et al., J. Pharmacol. Exp. Theor., 289, 1999.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 51<br />

effect of size, charge, and ligand modification of macromolecules on hepatic uptake clearance and<br />

urinary excretion clearance that essentially decide the pharmacokinetic properties of the macromolecule.<br />

Understanding the effect of the physicochemical properties of macromolecules on their<br />

tissue distribution allows theoretical design delivery systems, involving nanocarriers.<br />

4.3.1 EFFECT OF SIZE ON THE PHARMACOKINETIC PROPERTIES<br />

After intravenous injection, the kidney plays an important role in the disposition of macromolecules<br />

circulating in the blood. Macromolecules with a molecular weight of less than 50,000 (approximately<br />

6 nm in diameter) are susceptible to glomerular filtration and are excreted in the urine. 72–75<br />

Therefore, low molecular weight drugs and oligonucleotides with a lower molecular weight<br />

undergo rapid glomerular filtration <strong>by</strong> the kidney. 76,77 On the other hand, pDNA is too large to<br />

be filtered without degradation. 78<br />

4.3.2 EFFECT OF CHARGE ON THE PHARMACOKINETIC PROPERTIES<br />

Electric charge is also an important factor in determining the pharmacokinetic properties of macromolecules.<br />

79 Because the glomerular capillary walls also function as a charge-selective barrier<br />

having negative charges, positively charged macromolecules exhibit higher glomerular permeability<br />

than anionic macromolecules of similar molecular weights. Larger molecular weight<br />

cationic macromolecules escaping glomerular filtration mainly accumulate in the liver, kidney,<br />

and lung. 80–82 This phenomenon can be explained <strong>by</strong> following factors:<br />

1. Because the cell surface is negatively charged in general, cationic molecules tend to<br />

electrostatically bind to it.<br />

2. Because the liver and kidney have fenestrated capillaries, macromolecules can cross the<br />

endothelial barrier.<br />

3. Direct interaction with the lung endothelial cells.<br />

4. Embolization of aggregates of the complex with negatively charged blood components.<br />

On the other hand, strongly negatively charged macromolecules are taken up <strong>by</strong> liver nonparenchymal<br />

cells through liver fenestrae <strong>by</strong> scavenger receptor-mediated endocytosis. 78,83 Neutral<br />

or weakly anionic macromolecules with a molecular size that does not allow glomerular filtration<br />

such as polyethylene glycol and serum albumin posses a very weak affinity for the cell surface, and<br />

they are cleared very slowly. 7,8,84,85 Consequently, such neutral or weakly anionic macromolecules<br />

have a longer elimination half-life compared with the approximately same sized macromolecule<br />

with a strong negative or positive charge.<br />

4.3.3 EFFECT OF MODIFICATION FOR CELL-SELECTIVE DELIVERY<br />

Ligand modification for active targeting using receptor-mediated uptake markedly affects the<br />

pharmacokinetic properties of macromolecules. 4 For example, pharmacokinetic studies of<br />

receptor-mediated targeting to liver parenchymal cells via galactose recognition have been<br />

carried out based on the clearance concept. Galactose modification increased the hepatic uptake<br />

of macromolecules of different molecular weights. 3 The clearance was almost the same as the<br />

hepatic flow rate in mice, showing that it is almost identical to the theoretical maximum value.<br />

4.4 NANOCARRIERS FOR GENE DELIVERY<br />

In addition to traditional drugs, pDNA has recently also been used as a drug for cancer therapy; for<br />

example, antigen-encoding pDNA has a potential application in DNA vaccine therapy for cancer.<br />

Moreover, new techniques to inhibit target gene expression through transcriptional regulation<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


52<br />

without any changes in the functions of other genes using synthetic oligonucleotides such as<br />

antisense DNA, decoy oligonucleotide, and siRNA have been also considered as a novel anti-tumor<br />

chemotherapy. However, because pDNA and synthetic oligonucleotides are easily degraded or<br />

metabolized, there is no cell selectivity, and enough transgene expression or inhibition of target<br />

gene expression for effective therapy could not be achieved <strong>by</strong> the administration of naked pDNA<br />

or oligonucleotides. Therefore, for effective therapy, it is essential to develop a novel carrier that<br />

makes it possible to deliver such drugs to the target tissue or cells.<br />

4.4.1 POLYMERS<br />

Polymeric micelles prepared from cationic copolymer could be a novel gene carrier. After<br />

preparing a pDNA complex (standard size 5,000–7,000 bp), the mean particle of the complex is<br />

about 50–200 nm. pDNA entrapped in polymeric micelles is highly resistant to nuclease<br />

degradation. 28<br />

Dendrimers are also an attractive carrier for gene delivery because they can interact on an<br />

electrostatic charge basis with biologically relevant polyanions such as nucleic acids 86 and pDNA 87<br />

because their surfaces are covered with primary amino groups. Moreover, chemical or biological<br />

modification of the surface of dendrimers would make them more effective drug or gene carriers.<br />

4.4.2 LIPOSOMES<br />

Liposomes with cationic charge on their surface are novel carriers for genes. Several kinds of<br />

liposomes for gene delivery have been developed depending on passive or active targeting<br />

mechanisms. Table 4.2 summarizes liposomes for gene delivery.<br />

4.4.2.1 Cationic Liposomes for Lung Targeting<br />

After intravenous administration of pDNA complex, the immediate effect on the complex of<br />

erythrocytes is to induce aggregation. 88 Large aggregates (O400 nm) may be readily entrapped<br />

in the lung capillary. Mahato et al. demonstrated that cationic liposome/[ 32 P] pDNA complexes<br />

(lipoplexes) were rapidly cleared from the blood circulation and accumulated extensively in the<br />

lung and liver. 89 In addition, [ 32 P] lipoplexes were predominantly taken up <strong>by</strong> the liver NPC and<br />

TABLE 4.2<br />

In Vivo Targeted Gene Delivery<br />

Nanotechnology for Cancer Therapy<br />

System Liposome Target Tissue (Cell) Reference<br />

Capillary trapping Cationic-liposome Lung 90<br />

89<br />

91<br />

93<br />

Asialoglycoprotein receptor madiated Asialofetuin-liposome Liver (hepatocyte) 96<br />

Galactose-liposome Liver (hepatocyte) (hepatocyte) 97<br />

60<br />

98<br />

Mannose receptor mediated Mannose-liposome Liver (non-parenchymal cell) 66<br />

94<br />

95<br />

Folate receptor mediated Folate-liposome Tumor cell 99<br />

100<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 53<br />

this uptake was inhibited <strong>by</strong> the pre-administration of dextran sulfate, suggesting the involvement<br />

of a phagocytic process. However, gene expression after intravenous administration of lipoplexes<br />

was extremely high in the lung. 90–92 It is well known that IFNb exhibits anti-tumor activity. Taking<br />

this into consideration, intravenously injected pDNA encoding IFNb complexed with cationic<br />

liposome was found to significantly inhibit established metastatic lung tumors in CT-26 cells<br />

inoculated into mice. 93<br />

4.4.2.2 Mannose Liposomes for Liver NPC Targeting<br />

Because Man-C4-Chol has a cationic charge, a cationic charge could also be induced on the<br />

surface of a Man-liposome. Therefore, Man-liposomes are also an attractive potential gene carrier<br />

because of this cationic charge that would allow electrical interaction with the gene. As far as<br />

in vivo gene transfection is concerned, the highest gene expression was observed in the liver after<br />

intravenous injection of pDNA/Man-liposome complexes (Man-lipoplex) in mice. 66–68 In<br />

addition, gene transfer <strong>by</strong> a Man-liposome was mannose receptor-mediated because the gene<br />

expression with Man-lipoplex in the liver was significantly reduced <strong>by</strong> pre-dosing with mannosylated<br />

bovine serum albumin. Because macrophages in the liver and spleen exist at the<br />

endothelial cells intervals and could make contact with the complex without passing through<br />

the sinusoid (100–200 nm), cell-selective gene transfection could be achieved <strong>by</strong> intravenous<br />

administration of Man-lipoplex. Therefore, a mannosylated gene carrier would be effective for<br />

the NPC-selective gene transfection system even after intravenous administration. 94 As far as the<br />

application to cancer therapy using targeted gene delivery <strong>by</strong> mannosylated liposomes is<br />

concerned, DNA vaccine therapy is suitable because antigen encoded pDNA can be efficiently<br />

transfected into dendritic cells that expressed a large number of mannose receptors. Recently,<br />

Hattori et al. 95 demonstrated that the targeted delivery of a DNA vaccine <strong>by</strong> Man-liposomes is a<br />

potent method for DNA vaccine therapy.<br />

4.4.2.3 Galactose Liposomes for Liver PC Targeting<br />

PC exclusively expresses large numbers of high affinity cell-surface receptors that can bind asialoglycoprotein<br />

and internalize to the cell interior. In order to achieve PC-specific gene transfection,<br />

galactose moieties are introduced into the cationic liposomes. The galactosylation of liposomes can<br />

be achieved <strong>by</strong> coating with either glycoproteins or galactose conjugated synthetic lipids. Hara et<br />

al. 96 reported that asialofetuin-labeled liposomes encapsulating pDNA were taken up <strong>by</strong> asialoglycoprotein<br />

receptor-mediated endocytosis using cultured PC and showed the highest hepatic gene<br />

expression after intraportal injection with a preloading of EDTA. Gal-C4-Chol that possess a<br />

similar structure to Man-C4-Chol was synthesized. 58 In vivo gene transfer was examined <strong>by</strong><br />

optimizing the pharmacokinetics and physicochemical properties. 97 The radioactivity in the liver<br />

from the Gal-liposome/[ 32 P] pDNA complex (Gal-lipoplexes) was about 75% of the dose even<br />

1 min after intraportal administration. The hepatic gene expression of Gal-lipoplexes was more than<br />

a 10-fold greater than that of lipoplexes with bare cationic liposomes. When gene expression was<br />

examined at the intrahepatic cellular level, the gene expression of PC of Gal-lipoplexes was<br />

significantly higher than that of liver NPC. On the other hand, the gene expression of PC and<br />

NPC of lipoplexes was almost identical. In addition, asialoglycoprotein receptor-mediated endocytosis<br />

was also confirmed <strong>by</strong> the inhibitory effect of pre-dosing an excess amount of galactosylated<br />

bovine serum albumin. It was previously reported that the lipoplexes interact with erythrocytes after<br />

intravenous administration. 88 Recently, the presence of an essential amount of sodium chloride<br />

(NaCl) during the formation of Gal-lipoplexes stabilizing the complexes in accordance with the<br />

surface charge regulation (SCR) theory was demonstrated. 98 The transfection activity in hepatocytes<br />

of SCR Gal-lipoplexes was significantly higher than that of conventional complexes.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


54<br />

4.4.2.4 Folate Receptor-Mediated Targeting<br />

The folate receptor is known to be over expressed in a large fraction of human tumors, but it is only<br />

minimally distributed in normal tissues. Therefore, the folate receptor has also been used as a tumortargeting<br />

ligand for several drug delivery systems. Recently, Hofland et al. synthesized folate-PEGlipid<br />

derivatives to prepare folate-modified cationic liposomes. 99 After an intravenous injection of<br />

folate–conjugated liposome complex, gene expression in the tumors was not changed, whereas gene<br />

expression in the lung was reduced as compared with the conventional complex. However, after<br />

intraperitoneal injection in a murine disseminated peritoneal tumor model, folate–conjugated liposome<br />

complex formulations produced approximately a 10-fold increase in tumor-associated gene<br />

expression compared with conventional complex. 100<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Sezaki, H., Takakura, Y., and Hashida, M., Soluble macromolecular carriers for the delivery of<br />

antitumour drugs, Adv. Drug Deliv. Rev., 3, 247, 1989.<br />

2. Sezaki, H. and Hashida, M., Macromolecule–drug conjugates in targeted cancer chemotherapy, Crit.<br />

Rev. Theor. Drug Carrier Syst., 1, 1, 1984.<br />

3. Takakura, Y. and Hashida, M., Macromolecular carrier systems for targeted drug delivery: Pharmacokinetic<br />

considerations on biodistribution, Pharm. Res., 13, 820, 1996.<br />

4. Hashida, M., Kawakami, S., and Yamashita, F., Lipid carrier systems for targeted drug and gene<br />

delivery, Chem. Pharm. Bull., 53, 871, 2005.<br />

5. Takakura, Y. and Hashida, M., Macromolecular drug carrier systems in cancer chemotherapy:<br />

Macromolecular prodrugs, Crit. Rev. Oncol. Hematol., 18, 207, 1995.<br />

6. Jain, R. K., Transport of molecules across tumor vasculature, Cancer Metastasis Rev., 6, 559, 1987.<br />

7. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the anti-tumor agent<br />

smancs, Cancer Res., 46, 6387, 1986.<br />

8. Maeda, H. and Matsumura, Y., Tumoritropic and lymphotropic principles of macromolecular drugs,<br />

Crit. Rev. Theor. Drug Carrier Syst., 6, 193, 1989.<br />

9. Takakura, Y. et al., Influence of physicochemical properties on pharmacokinetics of non-viral<br />

vectors for gene delivery, J. Drug Target., 10, 99, 2002.<br />

10. Nishikawa, M. and Hashida, M., Pharmacokinetics of anti-cancer drugs, plasmid DNA, and their<br />

delivery systems in tissue-isolated perfused tumors, Adv. Drug Deliv. Rev., 40, 19, 1999.<br />

11. Ringsdorf, H., Structure and properties of pharmacologically active polymers, J. Polym. Sci. Polym.<br />

Symp., 51, 135, 1975.<br />

12. Duncan, R. et al., Polymers containing enzymatically degradable bonds 7. Design of oligopeptide<br />

sidechains in poly(N-(2-hydroxypropyl) metacylamide) copolymers to promote efficient degradation<br />

<strong>by</strong> lysosomal enzymes, Makromol. Chem., 184, 1997–2008, 1983.<br />

13. Kopecek, J. and Bazilova, H., Poly[N-(hydroxypropyl)-methacrylamide]—1. Radical polymerisation<br />

and copolymerization, Eur. Polym. J., 9, 7, 1973.<br />

14. Rihova, B. et al., Biocompatibility of N-(2-hydroxypropyl) methacrylamide copolymers containing<br />

adriamycin, Biomaterials, 10, 335, 1989.<br />

15. Cartlidge, S. A. et al., Soluble, crosslinked N-(2-hydroxypropyl) methacrylamide copolymers as<br />

potential drug carriers 2: Effect of molecular weight on blood clearance and body distribution in<br />

the rat after intravenous administration. Distribution of unfractionated copolymer after intraperitoneal,<br />

subcutaneous or oral administration, J. Control Release, 4, 253, 1987.<br />

16. Seymour, L. W. et al., Effect of molecular weight of N-(2-hydroxypropyl)methacrylamide copolymers<br />

on body distribution and rate of excretion after subcutaneous, intraperitoneal, and intravenous<br />

administration to rats, J. Biomed. Mater. Res., 21, 1341, 1987.<br />

17. Rejmanova, P. et al., Stability in rat plasma and serum of lysosomally degradable oligopeptide<br />

sequences in N-(2-hydroxypropyl)methacrylamide) copolymers, Biomaterials, 6, 45, 1985.<br />

18. Seymour, L. W. et al., Tumour tropism and anti-cancer efficacy of polymer-based doxorubicin<br />

prodrugs in the treatment of subcutaneous murine B16F10 melanoma, Br. J. Cancer, 70, 636, 1994.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 55<br />

19. Cassidy, J. et al., Clinical trials of nimodipine as a potential neuroprotector in ovarian cancer patients<br />

treated with cisplatin, Cancer Chemother. Pharmacol., 44, 161, 1998.<br />

20. Duncan, R. et al., Degradation of side-chains of N-(2-hydroxypropyl)methacrylamide) copolymers<br />

<strong>by</strong> lysosomal thiolproteinases, Biosci. Rep., 2, 1041, 1982.<br />

21. Seymour, L. W. et al., Pharmacokinetics of polymer-bound Adriamycin, Biochem. Pharmacol., 39,<br />

1125, 1990.<br />

22. Flanagan, P. A. et al., Evaluation of protein-N-(2-hydroxypropyl) methacrylamide copolymer conjugates<br />

as targetable drug carriers 2. Body distribution of conjugates containing transferrin, antitransferrin<br />

receptor antibody or anti-Thy 1.2 antibody and effectiveness of transferrin-containing<br />

daunomycin conjugates against mouse L 1210 leukaemia in vivo, J. Control Release, 18, 25, 1992.<br />

23. Wedge, S. R., Duncan, R., and Kopeckova, P., Comparison of the liver subcelluar distribution of free<br />

daunomycin and that bound to galactosamine targeted N-(2-hydroxypropyl)methacrylamide copolymers,<br />

following intravenous administration in the rat, Br. J. Cancer, 63, 546, 1991.<br />

24. Maeda, H., Sawa, T., and Konno, T., Mechanism of tumor-targeted delivery of macromolecular<br />

drugs, including the EPR effect in solid tumor and clinical overview of the prototype polymeric drug<br />

SMANCS, J. Control Release, 74, 47, 2001.<br />

25. Yokoyama, M., Block copolymers as drug carriers, Crit. Rev. Theor. Drug Carrier Syst., 9, 213,<br />

1992.<br />

26. Kataoka, K. et al., Block copolymer micelles as vehicles for drug delivery, J. Control Release, 24,<br />

119, 1993.<br />

27. Torchilin, V. P., Structure and design of polymeric surfactant-based drug delivery systems,<br />

J. Control Release, 73, 137, 2001.<br />

28. Kakizawa, Y. and Kataoka, K., Block copolymer micelles for delivery of gene and related<br />

compounds, Adv. Drug Deliv. Rev., 54, 203, 2002.<br />

29. Yokoyama, M. et al., Selective delivery of adriamycin to a solid tumor using a polymeric micelle<br />

carrier system, J. Drug Target., 7, 171, 1999.<br />

30. Yokoyama, M. et al., Toxicity and anti-tumor activity against solid tumors of micelle-forming<br />

polymeric anti-cancer drug and its extremely long circulation in blood, Cancer Res., 51, 3229, 1991.<br />

31. Kwon, G. et al., Biodistribution of micelle-forming polymer–drug conjugates, Pharm. Res., 10, 970,<br />

1993.<br />

32. Kwon, G. et al., Enhanced tumor accumulation and prolonged circulation times of micelle-forming<br />

poly (ethylene oxide–aspartate) block copolymer–adriamycin conjugates, J. Control Release, 29, 17,<br />

1994.<br />

33. Hamaguchi, T. et al., NK105, a paclitaxel-incorporating micellar nanoparticle formulation, can<br />

extend in vivo antitumour activity and reduce the neurotoxicity of paclitaxel, Br. J. Cancer, 92,<br />

1240, 2005.<br />

34. Kawakami, S. et al., Biodistribution characteristics of all-trans retinoic acid incorporated in liposomes<br />

and polymeric micelles following intravenous administration, J. Pharm. Sci., 94, 2606, 2005.<br />

35. Kohori, F. et al., Preparation and characterization of thermally responsive block copolymer micelles<br />

comprising poly(N-isopropylacrylamide-b-DL-lactide), J. Control Release, 55, 87, 1998.<br />

36. Chung, J. E. et al., Effect of molecular architecture of hydrophobically modified poly(N-isopropylacrylamide)<br />

on the formation of thermoresponsive core-shell micellar drug carriers, J. Control<br />

Release, 53, 119, 1998.<br />

37. Le Garrec, D. et al., Optimizing pH-responsive polymeric micelles for drug delivery in a cancer<br />

photodynamic therapy model, J. Drug Target., 10, 429, 2002.<br />

38. Tomalia, D. A. et al., A new class of polymers: Starburst-dendritic macromolecules, Polym. J., 17,<br />

117, 1985.<br />

39. Buhleier, E., Wehner, W., and Vogtle, F., Synthesis, 405, 155, 1978.<br />

40. Newkome, G. R. et al., Cascade molecules: A new approach to micelles, A[27]-arborol, J. Org.<br />

Chem., 50, 2003–2004, 1985.<br />

41. Svenson, S. and Tomalia, D. A., Dendrimers in biomedical applications-reflections on the field, Adv.<br />

Drug Deliv. Rev., 57, 2106, 2005.<br />

42. Roberts, J. C., Bhalgat, M. K., and Zera, R. T., Preliminary biological evaluation of polyamidoamine<br />

(PAMAM) Starburst dendrimers, J. Biomed. Mater. Res., 30, 53, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


56<br />

Nanotechnology for Cancer Therapy<br />

43. Nigavekar, S. S. et al., 3 H dendrimer nanoparticle organ/tumor distribution, Pharm. Res., 21, 476,<br />

2004.<br />

44. Padilla De Jesus, O. L. et al., Polyester dendritic systems for drug delivery applications: In vitro and<br />

in vivo evaluation, Bioconjug. Chem., 13, 453, 2002.<br />

45. Allen, T. M. and Chonn, A., Large unilamellar liposomes with low uptake into the reticuloendothelial<br />

system, FEBS Lett., 223, 42, 1987.<br />

46. Klibanov, A. L. et al., Amphipathic polyethyleneglycols effectively prolong the circulation time of<br />

liposomes, FEBS Lett., 268, 235, 1990.<br />

47. Managit, C. et al., Targeted and sustained drug delivery using PEGylated galactosylated liposomes,<br />

Int. J. Pharm., 266, 77, 2003.<br />

48. Maruyama, K., Kennel, S. J., and Huang, L., Lipid composition is important for highly efficient<br />

target binding and retention of immunoliposomes, Proc. Natl Acad. Sci. U.S.A, 87, 5744, 1990.<br />

49. Ashwell, G. and Harford, J., Carbohydrate-specific receptors of the liver, Annu. Rev. Biochem., 51,<br />

531, 1982.<br />

50. Kuiper, J., The Liver: Biology and Pthobiology, 3rd ed., Raven Press, New York, 1994.<br />

51. Weitman, S. D. et al., Distribution of the folate receptor GP38 in normal and malignant cell lines and<br />

tissues, Cancer Res., 52, 3396, 1992.<br />

52. Yuan, F. et al., Vascular permeability in a human tumor xenograft: Molecular size dependence and<br />

cutoff size, Cancer Res., 55, 3752, 1995.<br />

53. Gabizon, A. and Papahadjopoulos, D., Liposome formulations with prolonged circulation time in<br />

blood and enhanced uptake <strong>by</strong> tumors, Proc. Natl Acad. Sci. U.S.A, 85, 6949, 1988.<br />

54. Takino, T. et al., Long circulating emulsion carrier systems for highly lipophilic drugs, Biol. Pharm.<br />

Bull., 17, 121, 1994.<br />

55. Webb, M. S. et al., Sphingomyelin-cholesterol liposomes significantly enhance the pharmacokinetic<br />

and therapeutic properties of vincristine in murine and human tumour models, Br. J. Cancer, 72, 896,<br />

1995.<br />

56. Krishna, R. et al., Liposomal and nonliposomal drug pharmacokinetics after administration of liposome-encapsulated<br />

vincristine and their contribution to drug tissue distribution properties,<br />

J. Pharmacol. Exp. Theor., 298, 1206, 2001.<br />

57. Blume, G. and Cevc, G., Liposomes for the sustained drug release in vivo, Biochim. Biophys. Acta,<br />

1029, 91, 1990.<br />

58. Gabizon, A. A., Barenholz, Y., and Bialer, M., Prolongation of the circulation time of doxorubicin<br />

encapsulated in liposomes containing a polyethylene glycol-derivatized phospholipid: Pharmacokinetic<br />

studies in rodents and dogs, Pharm. Res., 10, 703, 1993.<br />

59. Gabizon, A. A., Pegylated liposomal doxorubicin: Metamorphosis of an old drug into a new form of<br />

chemotherapy, Cancer Invest., 19, 424, 2001.<br />

60. Kawakami, S. et al., Asialoglycoprotein receptor-mediated gene transfer using novel galactosylated<br />

cationic liposomes, Biochem. Biophys. Res. Commun., 252, 78, 1998.<br />

61. Kawakami, S. et al., Novel galactosylated liposomes for hepatocyte-selective targeting of lipophilic<br />

drugs, J. Pharm. Sci., 90, 105, 2001.<br />

62. Kawakami, S. et al., Biodistribution characteristics of mannosylated, fucosylated, and galactosylated<br />

liposomes in mice, Biochim. Biophys. Acta, 1524, 258, 2000.<br />

63. Kawakami, S. et al., Targeted delivery of prostaglandin E1 to hepatocytes using galactosylated<br />

liposomes, J. Drug Target., 8, 137, 2000.<br />

64. Hattori, Y. et al., Controlled biodistribution of galactosylated liposomes and incorporated probucol<br />

in hepatocyte-selective drug targeting, J. Control Release, 69, 369, 2000.<br />

65. Wang, S. N. et al., Synthesis of a novel galactosylated lipid and its application to the hepatocyteselective<br />

targeting of liposomal doxorubicin, Eur. J. Pharm. Biopharm., 62, 32, <strong>2006</strong>.<br />

66. Kawakami, S. et al., Mannose receptor-mediated gene transfer into macrophages using novel mannosylated<br />

cationic liposomes, Gene Theor., 7, 292, 2000.<br />

67. Kawakami, S. et al., The effect of lipid composition on receptor-mediated in vivo gene transfection<br />

using mannosylated cationic liposomes in mice, S.T.P. Pharm. Sci., 11, 117, 2001.<br />

68. Kawakami, S. et al., Effect of cationic charge on receptor-mediated transfection using mannosylated<br />

cationic liposome/plasmid DNA complexes following the intravenous administration in mice, Pharmazie,<br />

59, 405, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Pharmacokinetics of Nanocarrier-Mediated Drug and Gene Delivery 57<br />

69. Taniyama, T. and Holden, H. T., Direct augmentation of cytolytic activity of tumor-derived macrophages<br />

and macrophage cell lines <strong>by</strong> muramyl dipeptide, Cell. Immunol., 48, 369, 1979.<br />

70. Fox, A. and Fox, K., Rapid elimination of a synthetic adjuvant peptide from the circulation after<br />

systemic administration and absence of detectable natural muramyl peptides in normal serum at<br />

current analytical limits, Infect. Immunol., 59, 1202, 1991.<br />

71. Opanasopit, P. et al., Inhibition of liver metastasis <strong>by</strong> targeting of immunomodulators using mannosylated<br />

liposome carriers, J. Control Release, 80, 283, 2002.<br />

72. Brenner, B. M., Hostetter, T. H., and Humes, H. D., Glomerular permselectivity: Barrier function<br />

based on discrimination of molecular size and charge, Am. J. Physiol., 234, F455, 1978.<br />

73. Maack, T. et al., Renal filtration, transport, and metabolism of low-molecular-weight proteins: A<br />

review, Kidney Int., 16, 251, 1979.<br />

74. Mihara, K. et al., Disposition characteristics of protein drugs in the perfused rat kidney, Pharm. Res.,<br />

10, 823, 1993.<br />

75. Mihara, K. et al., Disposition characteristics of model macromolecules in the perfused rat kidney,<br />

Biol. Pharm. Bull., 16, 158, 1993.<br />

76. Sawai, K. et al., Renal disposition characteristics of oligonucleotides modified at terminal linkages in<br />

the perfused rat kidney, Antisense Res. Dev., 5, 279, 1995.<br />

77. Sawai, K. et al., Disposition of oligonucleotides in isolated perfused rat kidney: Involvement of<br />

scavenger receptors in their renal uptake, J. Pharmacol. Exp. Theor., 279, 284, 1996.<br />

78. Kawabata, K., Takakura, Y., and Hashida, M., The fate of plasmid DNA after intravenous injection<br />

in mice: Involvement of scavenger receptors in its hepatic uptake, Pharm. Res., 12, 825, 1995.<br />

79. Hashida, M. et al., Pharmacokinetics and targeted delivery of proteins and genes, J. Control Release,<br />

41, 91, 1996.<br />

80. Fujita, T. et al., Targeted delivery of human recombinant superoxide dismutase <strong>by</strong> chemical modification<br />

with mono- and polysaccharide derivatives, J. Pharmacol. Exp. Theor., 263, 971, 1992.<br />

81. Mahato, R. I. et al., In vivo disposition characteristics of plasmid DNA complexed with cationic<br />

liposomes, J. Drug Target., 3, 149, 1995.<br />

82. Ma, S. F. et al., Cationic charge-dependent hepatic delivery of amidated serum albumin, J. Control<br />

Release, 102, 583, 2005.<br />

83. Yamasaki, Y. et al., Pharmacokinetic analysis of in vivo disposition of succinylated proteins targeted<br />

to liver nonparenchymal cells via scavenger receptors: Importance of molecular size and negative<br />

charge density for in vivo recognition <strong>by</strong> receptors, J. Pharmacol. Exp. Theor., 301, 467, 2002.<br />

84. Fujita, T. et al., Control of in vivo fate of albumin derivatives utilizing combined chemical modification,<br />

J. Drug Target., 2, 157, 1994.<br />

85. Yabe, Y. et al., Targeted delivery and improved therapeutic potential of catalase <strong>by</strong> chemical<br />

modification: Combination with superoxide dismutase derivatives, J. Pharmacol. Exp. Theor.,<br />

289, 1176, 1999.<br />

86. Bielinska, A. et al., Regulation of in vitro gene expression using antisense oligonucleotides or<br />

antisense expression plasmids transfected using starburst PAMAM dendrimers, Nucleic Acids<br />

Res., 24, 2176, 1996.<br />

87. Tang, M. X., Redemann, C. T., and Szoka, F. C., In vitro gene delivery <strong>by</strong> degraded polyamidoamine<br />

dendrimers, Bioconjug. Chem., 7, 703, 1996.<br />

88. Sakurai, F. et al., Interaction between DNA–cationic liposome complexes and erythrocytes is an<br />

important factor in systemic gene transfer via the intravenous route in mice: The role of the neutral<br />

helper lipid, Gene Theor., 8, 677, 2001.<br />

89. Mahato, R. I. et al., Physicochemical and pharmacokinetic characteristics of plasmid DNA/cationic<br />

liposome complexes, J. Pharm. Sci., 84, 1267, 1995.<br />

90. Zhu, N. et al., Systemic gene expression after intravenous DNA delivery into adult mice, Science,<br />

261, 209, 1993.<br />

91. Song, Y. K. et al., Characterization of cationic liposome-mediated gene transfer in vivo <strong>by</strong> intravenous<br />

administration, Hum. Gene Theor., 8, 1585, 1997.<br />

92. Uyechi, L. S. et al., Mechanism of lipoplex gene delivery in mouse lung: Binding and internalization<br />

of fluorescent lipid and DNA components, Gene Theor., 8, 828, 2001.<br />

93. Sakurai, F. et al., Therapeutic effect of intravenous delivery of lipoplexes containing the interferonbeta<br />

gene and poly I: poly C in a murine lung metastasis model, Cancer Gene Theor., 10, 661, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


58<br />

Nanotechnology for Cancer Therapy<br />

94. Yamada, M. et al., Tissue and intrahepatic distribution and subcellular localization of a mannosylated<br />

lipoplex after intravenous administration in mice, J. Control Release, 98, 157, 2004.<br />

95. Hattori, Y. et al., Enhancement of immune responses <strong>by</strong> DNA vaccination through targeted gene<br />

delivery using mannosylated cationic liposome formulations following intravenous administration in<br />

mice, Biochem. Biophys. Res. Commun., 317, 992, 2004.<br />

96. Hara, T. et al., Receptor-mediated transfer of pSV2CAT DNA to a human hepatoblastoma cell line<br />

HepG2 using asialofetuin-labeled cationic liposomes, Gene, 159, 167, 1995.<br />

97. Kawakami, S. et al., In vivo gene delivery to the liver using novel galactosylated cationic liposomes,<br />

Pharm. Res., 17, 306, 2000.<br />

98. Fumoto, S. et al., Enhanced hepatocyte-selective in vivo gene expression <strong>by</strong> stabilized galactosylated<br />

liposome/plasmid DNA complex using sodium chloride for complex formation, Mol. Theor.,<br />

10, 719, 2004.<br />

99. Hofland, H. E. et al., Folate-targeted gene transfer in vivo, Mol. Theor., 5, 739, 2002.<br />

100. Reddy, J. A. et al., Folate-targeted, cationic liposome-mediated gene transfer into disseminated<br />

peritoneal tumors, Gene Theor., 9, 1542, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


5<br />

CONTENTS<br />

Multifunctional Nanoparticles<br />

for Cancer Therapy<br />

Todd J. Harris, Geoffrey von Maltzahn,<br />

and Sangeeta N. Bhatia<br />

5.1 Introduction ........................................................................................................................... 59<br />

5.2 Modular Functionalities at the Biosynthetic Interface ......................................................... 60<br />

5.2.1 Targeting.................................................................................................................... 60<br />

5.2.2 Imaging Agents ......................................................................................................... 61<br />

5.2.3 Sensing ...................................................................................................................... 64<br />

5.2.4 Therapeutic Payloads ................................................................................................ 66<br />

5.2.5 Remote Actuation...................................................................................................... 69<br />

5.3 Challenges in Integrating Multiple Functionalities and Future Directions ......................... 70<br />

References ...................................................................................................................................... 70<br />

5.1 INTRODUCTION<br />

The use of nanoparticles in cancer therapy is attractive for several reasons: they exhibit unique<br />

pharmacokinetics, including minimal renal filtration; they have high surface-to-volume ratios<br />

enabling modification with various surface functional groups that home, internalize, or stabilize;<br />

and they may be constructed from a wide range of materials used to encapsulate or solubilize<br />

therapeutic agents for drug delivery or to provide unique optical, magnetic, and electrical properties<br />

for imaging and remote actuation. The topology of a nanoparticle—core, coating, and surface<br />

functional groups—makes it particularly amenable to modular design, where<strong>by</strong> features and<br />

functional moieties may be interchanged or combined. Although many functionalities of nanoparticles<br />

have been demonstrated, including some clinically approved drug formulations and imaging<br />

agents, 3,8 the consolidation of these into multifunctional nanoparticles capable of targeting,<br />

imaging, and delivering therapeutics is an exciting area of research that holds great promise for<br />

cancer therapy in the future.<br />

Figure 5.1 1 schematically depicts a hypothetical multifunctional particle that has been engineered<br />

to include many features such as the ability to target tumors, evade uptake <strong>by</strong> the<br />

reticuloendothelial system (RES), protect therapeutics that can be released on demand, act as<br />

sensors of tumor responsiveness, and provide image contrast to visualize sites of disease and<br />

monitor disease progression. Some of these features, such as targeting, leverage biological<br />

machinery. Others are derived synthetically and enable external probing or manipulation that is<br />

otherwise not feasible in biological systems. In this chapter, we review both bio-inspired and<br />

synthetic nanoparticle functionalities that have been used in cancer therapy and address both<br />

current efforts and future opportunities to combine these into multifunctional devices.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

59


60<br />

Targeting species<br />

Therapeutic<br />

5.2 MODULAR FUNCTIONALITIES AT THE BIOSYNTHETIC INTERFACE<br />

5.2.1 TARGETING<br />

Diagnostic<br />

Actuate/Trigger<br />

Sensor<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 5.1 Schematic depiction of a multifunctional nanoparticle. A hypothetical nanoparticle targets the<br />

tumor, senses and reports molecular signatures, and delivers a therapeutic in response to an external or<br />

biological trigger. (From Ruoslahti, E., Cancer Cell, 2, 97–98, 2002.)<br />

The ability to physically target diseased cells to receive therapeutics while avoiding residual uptake<br />

in other tissues has long been a goal in cancer therapy. 9–12 The homing of stem cells to a tissue<br />

niche, or the susceptibility of one cell over another to viral infection, demonstrates that biomolecular<br />

recognition can be used to direct species to specific extracellular and intracellular sites. The<br />

microenvironment of the tumor including cell-surface markers, extracellular matrix, soluble<br />

factors, and proteases, as well as the tumor’s unique architecture and transport properties, may<br />

be exploited for targeting. 13–17<br />

Both passive and active targeting have been utilized for nanoparticle delivery. Passive targeting<br />

relies upon the unique pharmacokinetics of nanoparticles including minimal renal clearance and<br />

enhanced permeability and retention (EPR) through the porous angiogenic vessels in the tumor. 18,19<br />

Surface attachment of polymers such as poly(ethylene glycol) (PEG) and poly(ethylene oxide)<br />

(PEO) enables nanoparticles to avoid uptake <strong>by</strong> mononuclear phagocytes in the liver, spleen, and<br />

lymph nodes, there<strong>by</strong> improving accumulation in the tumor. 20–22 Active targeting relies on liganddirected<br />

binding of nanoparticles to receptors expressed in the tumor. Binding of ligands to the<br />

vasculature can occur immediately, as it is directly accessible to nanoparticles circulating in the<br />

blood. Over longer time periods, particles extravasate into the tissues where receptors expressed on<br />

cancer cells and in the interstitium may be used for localization. 23–25<br />

Many candidate tumor markers have been described, some of which bind known ligands such<br />

as arginine–glycine–aspartic acid (RGD)-binding Vb3 and Vb5 integrins expressed on the surface of<br />

angiogenic blood vessels, and folic acid-binding receptors on the surface of cancer cells. These and<br />

others have been attached to the surface of various nanoparticle cores to deliver them to<br />

tumors. 17,26–28 Monoclonal antibodies have also been used extensively for targeting. These can<br />

be isolated with high affinity for tumor markers and are useful for targeting receptors of unknown or<br />

low affinity ligands. 29,30 Novel screens for discovering tumor homing ligands have been developed<br />

using phage and bacterial display as well as libraries of aptamers, peptides, polymers, and small<br />

molecules. 31,32 These techniques may be used to isolate targeting ligands, even when their target<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Multifunctional Nanoparticles for Cancer Therapy 61<br />

receptor is unknown. For example, the 34 amino acid, cationic peptide F3, which has been used to<br />

deliver quantum dots to tumor endothelium, was uncovered initially <strong>by</strong> a blind-page display screen<br />

in a breast cancer xenograft model and later found to bind cell surface nucleolin expressed on tumor<br />

endothelium and cancer cells. 6,33,34<br />

Although extracellular targeting to the tumor is sufficient for many modes of imaging and drug<br />

delivery, intracellular delivery of nanoparticles into the cytosol is essential for some applications.<br />

For example, nanoparticles carrying membrane-impermeable cargo that perform their function in<br />

the cytosol, such as siRNA, antisense DNA, peptides, and other drugs, are minimally effective if<br />

delivered extracellularly or sequestered in the endosome. 35 Protein and peptide motifs capable of<br />

translocating nanoparticles into the cytoplasm have been borrowed from mechanisms of viral<br />

transfection. Two important classes of translocating domains include polycationic sequences and<br />

membrane fusion domains. Attaching the short polycationic sequence of HIV’s TAT protein, amino<br />

acid residues 48–57, to a nanoparticle facilitates its adsorption on a cell surface and subsequent<br />

internalization into the cell. 36,37 This peptide has been used to internalize dextran coated iron-oxide<br />

nanoparticles into T-cells in vitro, which were subsequently used to monitor T-cell trafficking in<br />

tumors with MRI. 38 Use of this peptide for intracellular delivery in vivo is limited <strong>by</strong> the adverse<br />

effect that polycationic sequences have on nanoparticle circulation time and RES uptake. 39 The<br />

amphiphilic domain derived from the N-terminus of the influenza protein hemagluttinin (HA2) is a<br />

membrane fusion peptide that destabilizes the endosome at low pH and facilitates viral escape into<br />

the cytosol. 40 Variations of this peptide with improved infectivity have also been synthesized. 41<br />

Influenza-derived peptides have been used to enhance the delivery of liposomes as well as 100 nm<br />

poly-L-lysine particles. Although the peptide modification of these particles improves endosomal<br />

escape over unmodified particles, the transfection efficiency still remains well below that of intact<br />

viruses. 42,43<br />

Another level of targeting can occur after translocation of nanoparticles into the cytosol to<br />

direct nanoparticles to specific sub-cellular structures. Using peptide localization sequences, fluorescent<br />

quantum dots have been targeted to the nucleus and the mitochondria (Figure 5.2). 2 Several<br />

other localization sequences exist and could be used to traffic nanoparticles to the endoplasmic<br />

reticulum, golgi apparatus, or peroxisomes. Although work in this area has been focused on<br />

organelle labeling, the potential for delivering therapeutic nanoparticles to sub-cellular structures<br />

is possible. Such nanoparticles could sense sub-cellular aspects of disease or specifically intervene<br />

for more potent treatment or eradication of cancer cells (i.e., free-radical-mediated mitochondrial<br />

damage to induce apoptosis).<br />

5.2.2 IMAGING AGENTS<br />

Imaging cancer is crucial for guiding decisions about treatment and for monitoring the efficacy of<br />

administered therapies. The use of nanoparticles for image contrast and enhancement has enabled<br />

improvements in cancer imaging <strong>by</strong> conventional modalities, such as magnetic resonance imaging<br />

(MRI) and ultrasound, and has also established new techniques such as optical-based imaging for<br />

cancer detection. 39,44,45 Targeted imaging agents that can identify specific biomarkers have the<br />

potential to improve detection, classification, and treatment of cancer with minimal invasiveness<br />

and reduced costs.<br />

The use of nanoparticles in cancer imaging has already demonstrated clinical efficacy in<br />

detecting liver cancer and staging lymph node metastasis noninvasively. 3,46 Superparamagnetic<br />

iron-oxide nanoparticles disrupt local magnetic field gradients in tissues, causing a detectable signal<br />

void in MRI. Dextran coated iron-oxide nanoparticles administered intravenously get phagocytosed<br />

<strong>by</strong> normal macrophages of the liver and lymph and the failure of these tissues to darken after ironoxide<br />

administration identifies invading cancer cells. Directly targeting these magnetic nanoparticles<br />

to cancer cells has also been demonstrated. For example, herceptin mAb and folic acid on the<br />

surface of iron-oxide nanoparticles enable MRI-based molecular imaging of their respective targets<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


62<br />

QD+NLS<br />

70kD Dextran Merge<br />

QD+MLS Mitotracker<br />

Merge<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 5.2 (See color insert following page 522.) Labeling of intracellular targets with peptide labeled<br />

quantum dots. Quantum dots (QD) modified with PEG and a nuclear localization sequence (NLS) or a<br />

mitochondrial localization sequence (MLS) were shown to target the nucleus or mitochondria of cells respectively.<br />

Seventy kilo Dalton PEG distributed in the cytoplasm contrasts nuclear localization while MitoTracker<br />

colocalizes with mitochondrial localization. (From Derfus, A. M., Chan, W. C. W., and Bhatia, S. N.,<br />

Advanced Materials, 16, 961, 2004.)<br />

in tumors. 47,48 Other nanoparticle cores including dendrimers, micelles, and liposomes modified<br />

with paramagnetic gadolinium have also been used for tumor targeted MRI contrast. 48–50<br />

Gold nanoshells offer a promising alternative to MRI probes <strong>by</strong> providing contrast for optical<br />

imaging. 45 These nanoparticles are constructed from a dielectric core (silicon) and a metallic<br />

conducting shell (gold). By varying the dimension of the core and shell, the plasmon resonance of<br />

these particles can be engineered to either absorb or scatter wavelengths of light, from UV to infrared.<br />

Particles that are tailored to scatter light in the near-infrared, where tissues have minimal absorbance,<br />

have been used to enhance imaging modalities such as reflectance confocal microscopy and optical<br />

coherence tomography (OCT). 51 Although the penetration of optical techniques does not approach<br />

that of CT or MRI, imaging features is possible at depths of a few centimeters. Gold colloids have also<br />

been used for optical contrast, but these lack the inherent tunability of nanoshells. The conjugation of<br />

optical contrast agents to antibodies has been used for the molecular imaging of the EGFR receptor on<br />

early cervical precancers and for Her2C breast carcinoma cells in mice. 52,53<br />

Fluorescent nanoparticles offer another useful tool to enhance optical detection. These probes<br />

are identified easily in microscopy and are useful for tracking the biodistribution of nanoparticles<br />

in experimental models. Fluorescent semiconductor nanocrystals, quantum dots, have been used<br />

to show ligand-mediated nanoparticle targeting to distinct features in the tumor. 6 Three different<br />

phage display-derived peptides were used to specifically target these nanocrystals to tumor blood<br />

vessels, tumor lymphatics, or lung endothelium (Figure 5.3). Functionalization of these nanoparticles<br />

with PEG eliminated detectable accumulation in RES organs, including the liver<br />

(Figure 5.4). Quantum dots have a distinct advantage over conventional fluorophores because<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Multifunctional Nanoparticles for Cancer Therapy 63<br />

(a)<br />

Prostate<br />

FIGURE 5.3 (See color insert following page 522.) Noninvasive detection of lymph node metastasis with<br />

iron-oxide nanoparticles and MRI. (a) A three-dimensional reconstruction using nanoparticle-enhanced MRI<br />

of the prostate. Metastatic lymph nodes are in red and normal nodes are in green. (b) Conventional MRI image<br />

shows similar signal intensity from two adjacent nodes. (c) Nanoparticle-enhanced MRI shows a decreased<br />

signal in the normal node from macrophage uptake (thick arrow), but not in the metastatic node (thin arrow).<br />

(From Harisinghani, M. G. and Weissleder, R., Plos Medicine, 1, 202–209, 2004.)<br />

of their size-tunable excitation and emission profiles, narrow bandwidths, and high photo-stability.<br />

54,55 Using nanocrystals that fluoresce in the near-infrared could extend their utility to clinical<br />

settings, 56 though a key limitation has been their potential toxicity because they are formulated<br />

from heavy metals. 57 Efforts to make these of nontoxic materials are ongoing (Figure 5.4).<br />

Alternative fluorescent nanoparticle probes have been developed including fluorescently tagged<br />

dendrimers and fluorophore-embedded silica nanoparticles. 58–60<br />

Nanoparticle formulations that provide contrast for other imaging modalities including ultrasound<br />

and CT have been described. Perfluorocarbon emulsion nanoparticles composed of lipidencapsulated<br />

perfluorocarbon liquid, about 250 nm diameter, are effective in giving echo contrast. 61<br />

Air-entrapping liposomes formulated from freeze-drying techniques have also been developed to<br />

give ultrasound contrast. 62,63 These agents passively distribute in RES organs and areas of angiogenenis<br />

enabling enhanced imaging of these features. Bismuth sulfide nanoparticles can be used as<br />

contrast agents for CT imaging, giving blood pool contrast similar to that of iodine but at lower<br />

concentrations. These can also be used to image lymph nodes after phagocytic uptake. 64 As with<br />

other imaging modalities, the attachment of appropriate ligands to nanoparticle-based CT and<br />

ultrasound contrast agents could be used for molecular imaging.<br />

An interesting extension of the imaging agents described above is their combination into<br />

multimodal imaging nanoparticles. The combination of fluorescence and magnetic properties<br />

within a single particle has been used for dual optical and MRI imaging. Fluorescence is used to<br />

determine the localization of targeted imaging agents down to specific micro-structures inside and<br />

outside cells and magnetic domains provide three-dimensional whole-body imaging capabilities<br />

with MRI. 59,65 These magneto-fluorescent nanoparticles could be effective for image guided<br />

surgeries where MRI is used to locate cancer in the body and fluorescence is used to more precisely<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

(b)<br />

(c)


64<br />

(a)<br />

(b)<br />

Peptide<br />

Peptide<br />

Peptide<br />

delineate tumor borders during resection. Other dual-imaging probes have been described<br />

including: perfluorocarbon emulsions tagged with gadolinium for combined ultrasound and MRI. 66<br />

5.2.3 SENSING<br />

Peptide<br />

ZnS<br />

CdSe<br />

Peptide<br />

Peptide<br />

Peptide<br />

Peptide<br />

Peptide<br />

Peptide<br />

PEG<br />

ZnS<br />

PEG<br />

Functionalities that undergo chemical alterations in response to enzymatic activity or other<br />

properties such as pH or oxygen could be used as sensors to report information about the status of<br />

the tumor or efficacy of treatment. Many nanoparticle-based sensors that respond to biological<br />

triggers including proteases, DNAses, proteins, peroxidase, pH, and others have been demonstrated<br />

in vitro. 67–72 These generally rely on assembly or disassembly of inorganic nanocrystals<br />

including: gold nanoparticles, nanoshells, or nanorods, which undergo a shift in their plasmon<br />

resonance when aggregated; iron-oxide nanoparticles have enhanced T2 relaxivity when clustered;<br />

and fluorescent quencher-based nanoparticle systems that dequench after triggered release.<br />

A system using cleavable polymeric shielding of self-assembling nanoparticles has been<br />

proposed as a mechanism for translating nanoparticle-based enzyme sensors to in vivo use<br />

(Figure 5.5). 73 Self-assembling, complementary iron-oxide nanoparticles are rendered latent with<br />

PEG polymers linked to the nanoparticle surface <strong>by</strong> protease-cleavable substrates that serve to both<br />

inhibit assembly and stabilize the particles in serum. Upon proteolytic removal of PEG polymers <strong>by</strong><br />

MMP-2 expressing cancer cells, nanoparticles assemble and acquire amplified magnetic properties<br />

that can be detected with MRI. In the future, similar to thrombin-driven self-assembly of fibrin and<br />

platelets at sites of endothelial injury, this system may allow the hyper-active proteolytic processes of<br />

cancers to drive the self-assembly of nanoparticles in regions of cancer angiogenesis, invasion, and<br />

metastasis in vivo. Due to its modular design, this system can easily be modified for a number of<br />

detection schemes <strong>by</strong> substituting the complementary binding pairs, cleavable substrates<br />

Peptide<br />

PEG CdSe Peptide<br />

FIGURE 5.4 (See color insert following page 522.) Targeting quantum dots (QDs) to site-specific endothelium<br />

with phage display-derived peptides. (a) Schematic representation of co-injected red and green<br />

quantum dots that home to tumor and lung vasculature respectively after intravenous injection. (b) Schematic<br />

representation of peptide-coated QDs and peptide-coated PEG-QDs. (c) QDs labeled with the tumor endothelium<br />

homing peptide F3 co-localize with a blood vessel marker. (d) QDs labeled with the tumor lymphatic<br />

homing peptide Lyp-1 highlight the endothelium but do not colocalize with a blood vessel marker. (From<br />

Akerman, M. E., Chan, W. C. W., Laakkonen, P., Bhatia, S. N., and Ruoslahti, E., Proceedings of the National<br />

Academy of Sciences of the United States of America, 99, 12617–12621, 2002.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

PEG<br />

(c)<br />

(d)<br />

Nanotechnology for Cancer Therapy


(a)<br />

(b)<br />

0.03<br />

0.02<br />

Δe<br />

0.01<br />

0.00<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

0<br />

No MMP-2<br />

Scmbl cntrl<br />

MMP-2<br />

30 60<br />

Diffusivity<br />

'Latent' nanoparticles Proteolytically actuated self-assembly<br />

90<br />

t /min<br />

120 150 180<br />

(c) (d)<br />

10 kDa PEG<br />

MMP-2<br />

MMP-2<br />

Nanoassemblies exhibit:<br />

Magnetic susceptibility<br />

T2 relaxativity<br />

No MMP-2<br />

32<br />

particle 10<br />

concn./pM<br />

3.2<br />

0<br />

85 170 340 680 1360<br />

MMP-2/ng/ml<br />

400<br />

300<br />

200<br />

T2/ms<br />

FIGURE 5.5 (See color insert following page 522.) Protease-triggered self-assembling nanoparticles with polymer-shielded coatings. (a) Schematic representation of<br />

nanoparticles that self-assemble after protease-mediated cleavage of PEG chains reveals complementary moieties (neutravidin and biotin). (b) Iron-oxide nanoparticles<br />

with cleavable linkers assemble in the presence of MMP-2, as measured <strong>by</strong> changes their light extinction, while particles with noncleavable scrambled peptides do not. (c)<br />

Atomic force micrographs of particles incubated with MMP-2 shows detectable aggregation (scale bars are 500 nm). (d) T2 maps of particles in solution using a 4.7 T<br />

MRI demonstrate enhanced T2 relaxivity for increasing concentrations of MMP-2.<br />

100<br />

Multifunctional Nanoparticles for Cancer Therapy 65


66<br />

(e.g., glycans, lipids, oligonucleotides), or multivalent nanoparticle cores (e.g., gold, quantum dot,<br />

dendrimer).<br />

5.2.4 THERAPEUTIC PAYLOADS<br />

Nanotechnology for Cancer Therapy<br />

The use of nanoparticulate drug carriers can address many critical challenges in drug delivery<br />

including: improving drug solubility and stability; extending drug half-lives in the blood; reducing<br />

adverse effects in nontarget organs; and concentrating drugs at the disease site. 74 Drugs may be<br />

dispersed in a matrix, encapsulated in a vesicle, dissolved in a hydrophobic core, or attached to the<br />

surface of a nanoparticle. Several nanoparticle-based drug delivery systems including liposomes,<br />

polymeric nanoparticles, dendrimers, ceramic based carriers, micelles, and others have been used to<br />

carry small molecule, peptide, and oligonucleotide therapeutic agents. 24,75 Many promising<br />

anti-cancer drugs fail to make it to the clinic because of poor solubility or high collateral toxicity<br />

at therapeutic levels, thus motivating the need for these carriers in cancer therapy.<br />

Liposomes have been the most extensively utilized nanoparticle-based carriers for delivering<br />

anti-cancer drugs. First described decades ago, these submicron-sized carriers consist of amphiphilic<br />

lipids assembled to form vesicles that can encapsulate drugs. 76 Liposome-encapsulated doxorubicin<br />

is a clinically approved nanoparticle formulation used for chemotherapy. 77 The surface of this<br />

nanocarrier is PEGylated to reduce rapid uptake <strong>by</strong> phagocytic cells and extend the drug circulation<br />

time for better therapeutic efficacy. Several other liposome-encapsulated chemotherapeutic drugs<br />

have been described, with many in clinical trials. 78 Active targeting of these liposomes through the<br />

attachment of antibodies and various ligands has also been demonstrated. 30,79 Drug loaded liposomes<br />

with encapsulated or surface-functionalized gadolinium or fluorophores have been used to simultaneously<br />

image tumors during nanoparticle-targeted drug delivery. 80–82<br />

Biodegradable polymer nanocarriers have also been investigated as a means of encapsulating<br />

drugs and releasing them over time. Both poly dl-lactide co-glycolide (PLGA) and polylactide (PLA)<br />

nanoparticles have been formed that immobilize drugs dispersed in their matrix and release them<br />

upon degradation. 24,83 Other polymers, including polyethyleneimine (PEI), polylysine, and cyclodextrin-containing<br />

polymers, are used to condense DNA or RNA into nanoparticle carriers that can<br />

be targeted to cancer cells for gene or siRNA delivery. 35 Polymeric micelles consist of amphiphilic<br />

block copolymers that self-assemble into a water-soluble nanoparticle with a hydrophobic core.<br />

These can be used to encapsulate water-insoluble drugs such as doxorubicin and adriamycin and<br />

targeted to tumors. 84–86 Polymersomes are another variation of polymer-based nanoparticulate<br />

vesicles that self-assemble from amphiphilic block copolymers. 87 These have been used to encapsulate<br />

doxorubicin with well-controlled release over several days. 88,89<br />

Another class of nanoparticle-based drug carriers are dendrimers. These consist of a network of<br />

branching chemical bonds around an inner core. One of the more popular dendrimers, polyamidoamine<br />

dendrimers (PAMAMs), are nonimmunogenic, water-soluble, and possess terminal amine<br />

functional groups for conjugation of a variety of surface moieties. 90 Their inner core can been used<br />

to encapsulate anti-cancer drugs such as doxorubicin and methotrexate. 91 Drugsmayalsobe<br />

conjugated to the dendrimer surface along with ligands for targeting. 92,93 A dendrimer functionalized<br />

with FITC, folic acid, and methoxetrate has been synthesized to have imaging, targeting, and<br />

drug delivery capabilities (Figure 5.6). 4,7,94 The synthesis of these conjugates in a scalable and<br />

reproducible manner has been described for potential clinical applications. 4<br />

Other nanoparticulate carriers, including nanoemulsions, drug nanocrystals, and polyelectrolyte<br />

carriers, have been developed. Nanoemulsions are formed <strong>by</strong> dissolving a drug in a lipid,<br />

cooling the solution under high pressure, and using homogenization to form solid nanoparticle lipid<br />

carriers at body temperature. Homogenization techniques can also be used to form crystalline<br />

nanosuspensions of drugs. 74 These formulations increase drug solubility and control release<br />

kinetics of the drug in the blood and at the tumor site. Polyelectrolyte carriers formed <strong>by</strong> the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


H 2 N<br />

N<br />

N<br />

NH 2<br />

H<br />

CH 3<br />

C<br />

HN<br />

O<br />

5<br />

HO CH O<br />

C<br />

O<br />

MTX<br />

O HO<br />

HO<br />

HO<br />

HO HOHO<br />

HO<br />

OH<br />

N<br />

N<br />

15<br />

O<br />

O<br />

C<br />

OH<br />

OH<br />

H<br />

N C<br />

N<br />

H<br />

O<br />

NH<br />

G5<br />

N<br />

H<br />

O<br />

C<br />

CH3 AC<br />

82<br />

HN<br />

C S<br />

HO<br />

HN<br />

O<br />

COOH<br />

4<br />

FITC<br />

(a) (b)<br />

FA<br />

O<br />

H<br />

H<br />

OH<br />

NH 2<br />

4<br />

G5-FI<br />

G5-FI-PA<br />

FIGURE 5.6 Multifunctional dendrimers for targeting, imaging, and drug delivery. (a) Schematic of a multifunctional dendrimer labeled with fluorescein (FITC) for<br />

imaging, folic acid (FA) for targeting, methotrexate (MTX) for therapeutic delivery, and alcohol (OH) and acetylated (Ac) moieties for particle stabilization. (b) Confocal<br />

images of FITC-labeled dendrimers incubated over cells with and without a targeting antibody, anti-PSMA (PA). (From Majoros, I. J., Thomas, T. P., Mehta, C. B., and<br />

Baker, J. R., Journal of Medicinal Chemistry, 48, 5892–5899, 2005; Thomas, T. P. et al., Biomacromolecules, 5, 2269–2274, 2004.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

1<br />

2<br />

Multifunctional Nanoparticles for Cancer Therapy 67


68<br />

layer-<strong>by</strong>-layer absorption of polycationic and polyanionic moieties can be used to encapsulate<br />

therapeutic cargo, particularly larger agents such as peptides and oligonucleotides. 95<br />

A clever combination of drug-release modalities was recently demonstrated <strong>by</strong> the creation of<br />

a dual drug-release nanoparticle having a PLGA polymer core encapsulating doxorubicin and a<br />

PEG-lipid block copolymer shell loaded with the combretastatin (Figure 5.7). 5 The lipophilic<br />

anti-angiogenesis drug, combretastatin, intercalates in the nanoparticle membrane and releases<br />

rapidly upon association with tumor endothelial cells, while the slower-releasing doxorubicin<br />

increases cytotoxic killing of tumor cells for a prolonged time after the vasculature shuts down.<br />

This novel system demonstrates the feasibility of integrating multiple functionalities of drug<br />

delivery on a single nanoparticle to enhance therapeutic efficacy.<br />

Total drug released<br />

(a) (c)<br />

(b) (d)<br />

Tumor volume<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 20 40 60 80 100 120<br />

Time (hours)<br />

7500<br />

5000<br />

2500<br />

Nanotechnology for Cancer Therapy<br />

Veh<br />

C<br />

N<br />

N+C<br />

NC<br />

0<br />

9 10 11 12 13<br />

Days<br />

Combretastatin (×10 2 μg)<br />

Doxorubicin (μg)<br />

140 160<br />

FIGURE 5.7 Dual drug-release nanoparticle for combined anti-angiogenisis and anti-cancer treatment. (a)<br />

Scanning electron micrograph showing nanocores prepared from doxorubicin-coupled PLGA. The nanocores<br />

are encapsulated inside a lipid coat, which is also loaded with an anti-angiogenesis agent. (b) Cross section of a<br />

nanocell with the dark nanocore. The lipid coat is surface-modified through pegylation, which confers stealth<br />

characteristics to the nanocell from the RES. (c) The composition of the nanocell enables a spatiotemporal<br />

release of the two agents in an acidic pH mimicking the tumor environment, as shown in the graph. (d) In vivo<br />

studies using F10 melanoma clearly show that the spatiotemporal release from the nanocells (NC) achieves<br />

better outcome than the doxorubicin-loaded nanocore or the lipid-entrapped combretastatin (C) alone, or<br />

combinations of both (N+C). 5<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

14


Multifunctional Nanoparticles for Cancer Therapy 69<br />

5.2.5 REMOTE ACTUATION<br />

Temporal and spatial control of therapeutic administration is important for eliminating off-target<br />

toxicity and achieving optimal delivery. Temporally controlled release profiles may be designed into<br />

nanoparticle carriers mentioned previously and spatial control can be improved with targeting.<br />

However, off-target effects, including eventual accumulation of nanoparticles in RES organs, limit<br />

many aspects of these methods of control. The ability to trigger the therapeutic activity of administered<br />

nanoparticles remotely could be a valuable tool for localizing treatments to a diseased site.<br />

Many inorganic nanocrystals and nanoemulsions used for imaging contrast absorb electromagnetic<br />

or ultrasonic energy that can also be used to remotely heat or trigger drug delivery.<br />

Thermal ablation of tumors <strong>by</strong> nanoparticles that absorb external energy has been demonstrated<br />

both with iron-oxide nanoparticles and gold nanoshells. Superparamagnetic iron-oxide nanoparticles<br />

under the influence of an alternating electromagnetic (EM) field heat <strong>by</strong> Brownian relaxation,<br />

where heat is generated <strong>by</strong> the rotation of particles in the field, and Neel relaxation, where the<br />

magnetic domains are moved away from their easy axis with the resultant energy being deposited as<br />

heat in the solution. 96,97 Nanoparticle concentrations of 0.1–1% are required to achieve critical<br />

temperatures for tumor ablation. 98,99 Ongoing work to increase the absorption of magnetic nanoparticles<br />

using clinically safe RF frequencies and to increase the concentration of particles that can<br />

be targeted to the tumor may extend the utility of this technique. Alternatively, near-infraredabsorbing<br />

gold nanoshells targeted to the tumor can be used to thermally ablate the cancer<br />

cells upon illumination with a high intensity laser. 100,101 This technique can be applied to solid<br />

tumors in close proximity to the skin, but cannot be applied to deeper lesions because of tissue<br />

absorbance. 53,100,101 By synthesizing nanoshells with a plasmon resonance that has both absorption<br />

and scattering profiles, these nanoparticles may be capable of both heating and imaging<br />

tumors. 53<br />

Remotely-triggered release of a therapy <strong>by</strong> heating is a promising extension of the use of<br />

nanoparticles that can absorb external energy. An example of this has been demonstrated with a<br />

model drug linked to an iron-oxide nanoparticle via a heat-labile tether that is released and diffuses<br />

into the peripheral tissue after irradiation with RF energy. 98 By modifying the susceptibility of the<br />

linker, it is possible to tune the release profile over a range of temperatures and to enable repeated<br />

administrations. The iron core of these drug-releasing nanoparticles can be used simultaneously for<br />

imaging with MRI. Additionally, the magnetic properties of these nanoparticles can be manipulated<br />

<strong>by</strong> magnetic field gradients to target sites near externally- or internally-placed magnets. 102<br />

Drug activation using EM energy has been explored extensively with photodynamic therapy<br />

(PDT). PDT agents, when irradiated <strong>by</strong> light, produce reactive oxygen species that are toxic to cells.<br />

Agents such as porphyrins have been conjugated to various nanoparticle cores including dendrimers,<br />

liposomes, and polymers. 103,104 When excited <strong>by</strong> light, these nanoparticles can produce<br />

enough reactive oxygen species to kill tumor cells. 60 The inherent fluorescent properties of<br />

many PDT agents enable simultaneous imaging with therapeutic delivery. A multifunctional nanoparticle<br />

platform combining MRI contrast and photodynamic therapy has been used to target,<br />

image, and treat brain cancer in a rat model. 105 In the future, integrating these nanoparticles<br />

with peptides capable of targeting tumors and subcellularly localizing them to the nuclei or mitochondria<br />

of tumor cells may enhance the therapeutic efficacy of these treatments.<br />

Other forms of externally applied energy such as ultrasound and x-ray radiation provide<br />

alternative mechanisms to achieve remote actuation. Acoustic energy has been shown to<br />

enhance the delivery of lipid drugs from a perfluorocarbon emulsion targeted to cell membranes<br />

and from doxorubicin-loaded polymeric micelles. 106,107 Atomically dense nanoparticles have been<br />

shown to increase the absorption of x-ray radiation, enhancing their destructive effect in<br />

surrounding tissue. 108 There is potential for simultaneous imaging and therapeutic delivery with<br />

these particles also.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


70<br />

5.3 CHALLENGES IN INTEGRATING MULTIPLE FUNCTIONALITIES AND<br />

FUTURE DIRECTIONS<br />

Although remotely actuated nanoparticle cores such as iron-oxide and metal nanoshells naturally<br />

lend themselves to dual-imaging and therapeutic applications, the combination of imaging and<br />

other functionalities using other nanoparticle cores can be challenging. There are inherent trade-offs<br />

when combining many functional groups into one nanoparticle. In many cases, a limited number of<br />

attachment sites are available on the particle surface, making it difficult to couple several functional<br />

groups in sufficient concentration for each to function. Moreover, some groups may interact to<br />

sterically shield or alter the activity of one another when combined in close proximity. Multiple<br />

functional moieties on a nanoparticle may also reduce colloidal stability or adversely affect its<br />

in vivo pharmacokinetics. With significant characterization and fine tuning, dendrimers that<br />

combine targeting, imaging, and therapeutic moieties on their surface have been synthesized<br />

successfully. 4 Similar efforts will be necessary to achieve other multifunctional nanoparticles<br />

with decorated-surface moieties.<br />

An alternative strategy to consolidate multiple functionalities onto a single particle is to use<br />

core-shell architecture. In this case, an outer shell with one functionality, such as targeting, may be<br />

unveiled to reveal an inner core that performs a secondary function such as endosomal escape or<br />

drug release. This has been demonstrated with the conjugation of targeting moieties or protective<br />

PEG groups on the surface of dendrimers or polymers via acid-labile chemistries that degrade in the<br />

lower pH of the endosome and unveil endosomal escape mechanisms on the particle core. 109,110<br />

This has also been demonstrated with protease-cleavable linkers that release protective polymers on<br />

the surface of complementary nanoparticles to initiate their self-assembly. 73<br />

The synthesis of nanoparticles with polar domains is another strategy that could be used to<br />

incorporate multiple functionalities on a single particle. Janus nanoparticles—named for Janus, the<br />

Roman God of doorways typically depicted with faces on the front and back of his head—have been<br />

engineered with two chemically distinct hemispheres or surfaces. These nanoparticles may be<br />

spherical (with opposing faces of unique composition), dumbbell-shaped (with two equal-sized<br />

spheres linked together), snowman-shaped, and may have other morphologies as well. 95,111 The<br />

creation of nanoparticles with spatially separated chemical domains is a step towards replicating the<br />

controlled polarity exhibited in nature across many length scales. Separate hemispheres may be<br />

used to isolate and organize functional domains on nanoparticles such that they may simultaneously<br />

carry targeting molecules, endosomal escape domains, sensing moieties, hydrophilic and hydrophobic<br />

therapeutics, or contrast agents that otherwise might be mutually inhibitory if randomly<br />

incorporated. Moreover, there may be specific applications for which the polarity and anisotropy of<br />

Janus nanoparticles have benefit, such as real-time detection of oriented binding events, targeted<br />

bridging of multiple components at a tumor cell, directed drug delivery, or guided self-assembly.<br />

Although there have been many exciting advances in the application of nanoparticles for cancer<br />

imaging and treatment, the true power of these materials will be in their ability to interact with<br />

disease processes intelligently. The modular design of functionalities that target, sense, signal, and<br />

treat and the ongoing efforts to consolidate these into single nanoparticle platforms is one way in<br />

which such ‘smart’ materials are being developed. The further elucidation of complex biological<br />

processes in tumorogenesis, the discovery of nanomaterials with other novel properties, and the<br />

consolidation of biological and synthetic machinery in these materials in new and elegant ways are<br />

key factors that will determine their future success in cancer therapy.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Ruoslahti, E., Antiangiogenics meet nanotechnology, Cancer Cell, 2, 97–98, 2002.<br />

2. Derfus, A. M., Chan, W. C. W., and Bhatia, S. N., Intracellular delivery of quantum dots for live cell<br />

labeling and organelle tracking, Advanced Materials, 16, 961–966, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Multifunctional Nanoparticles for Cancer Therapy 71<br />

3. Harisinghani, M. G. and Weissleder, R., Sensitive, noninvasive detection of lymph node metastases,<br />

Plos Medicine, 1, 202–209, 2004.<br />

4. Majoros, I. J., Thomas, T. P., Mehta, C. B., and Baker, J. R., Poly(amidoamine) dendrimer-based<br />

multifunctional engineered nanodevice for cancer therapy, Journal of Medicinal Chemistry, 48,<br />

5892–5899, 2005.<br />

5. Sengupta, S. et al., Temporal targeting of tumour cells and neovasculature with a nanoscale delivery<br />

system, Nature, 436, 568–572, 2005.<br />

6. Akerman, M. E., Chan, W. C. W., Laakkonen, P., Bhatia, S. N., and Ruoslahti, E., Nanocrystal<br />

targeting in vivo, Proceedings of the National Academy of Sciences of the United States of America,<br />

99, 12617–12621, 2002.<br />

7. Thomas, T. P. et al., In vitro targeting of synthesized antibody-conjugated dendrimer nanoparticles,<br />

Biomacromolecules, 5, 2269–2274, 2004.<br />

8. Gordon, A. N. et al., Recurrent epithelial ovarian carcinoma: A randomized phase III study of pegylated<br />

liposomal doxorubicin versus topotecan, Journal of Clinical Oncology, 19, 3312–3322, 2001.<br />

9. Ruoslahti, E., Drug targeting to specific vascular sites, Drug Discovery Today, 7, 1138–1143, 2002.<br />

10. Allen, T. M., Charrois, G. J. R., and Sapra, P., Recent advances in passively and actively targeted<br />

liposomal drug delivery systems for the treatment of cancer, Abstracts of Papers of the American<br />

Chemical Society, 226, U458, 2003.<br />

11. Allen, T. M. and Cullis, P. R., Drug delivery systems: Entering the mainstream, Science, 303,<br />

1818–1822, 2004.<br />

12. Wickham, T. J., Targeting adenovirus, Gene Therapy, 7, 110–114, 2000.<br />

13. Ruoslahti, E. and Rajotte, D., An address system in the vasculature of normal tissues and tumors,<br />

Annual Review of Immunology, 18, 813–827, 2000.<br />

14. Jain, R. K., Delivery of molecular and cellular medicine to solid tumors, Advanced Drug Delivery<br />

Reviews, 46, 149–168, 2001.<br />

15. Jain, R. K., Delivery of molecular and cellular medicine to solid tumors, Journal of Controlled<br />

Release, 53, 49–67, 1998.<br />

16. Jain, R. K., Delivery of molecular and cellular medicine to solid tumors, Microcirculation—London,<br />

4, 3–23, 1997.<br />

17. Satchi-Fainaro, R. et al., Targeting angiogenesis with a conjugate of HPMA copolymer and<br />

TNP-470, Nature Medicine, 10, 255–261, 2004.<br />

18. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancerchemotherapy—mechanism<br />

of tumoritropic accumulation of proteins and the antitumor agent<br />

Smancs, Cancer Research, 46, 6387–6392, 1986.<br />

19. Tabata, T., Murakami, Y., and Ikada, Y., Tumor accumulation of poly(vinyl alcohol) of different<br />

sizes after intravenous injection, Journal of Controlled Release, 50, 123–133, 1998.<br />

20. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

Theory to practice, Pharmacological Reviews, 53, 283–318, 2001.<br />

21. Moghimi, S. M. and Hunter, A. C., Recognition <strong>by</strong> macrophages and liver cells of opsonized<br />

phospholipid vesicles and phospholipid headgroups, Pharmaceutical Research, 18, 1–8, 2001.<br />

22. Nicolazzi, C. et al., Anionic polyethyleneglycol lipids added to cationic lipoplexes increase their<br />

plasmatic circulation time, Journal of Controlled Release, 88, 429–443, 2003.<br />

23. Ruoslahti, E., Specialization of tumour vasculature, Nature Reviews Cancer, 2, 83–90, 2002.<br />

24. Sahoo, S. K. and Labhasetwar, V., Nanotech approaches to delivery and imaging drug, Drug<br />

Discovery Today, 8, 1112–1120, 2003.<br />

25. Wickline, S. A. and Lanza, G. M., Nanotechnology for molecular imaging and targeted therapy,<br />

Circulation, 107, 1092–1095, 2003.<br />

26. Stella, B. et al., Design of folic acid-conjugated nanoparticles for drug targeting, Journal of Pharmaceutical<br />

Sciences, 89, 1452–1464, 2000.<br />

27. Lockman, P. R. et al., Brain uptake of thiamine-coated nanoparticles, Journal of Controlled Release,<br />

93, 271–282, 2003.<br />

28. Dubey, P. K., Mishra, V., Jain, S., Mahor, S., and Vyas, S. P., Liposomes modified with cyclic RGD<br />

peptide for tumor targeting, Journal of Drug Targeting, 12, 257–264, 2004.<br />

29. Kirpotin, D. et al., Sterically stabilized Anti-HER2 immunoliposomes: Design and targeting to<br />

human breast cancer cells in vitro, Biochemistry, 36, 66–75, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


72<br />

Nanotechnology for Cancer Therapy<br />

30. Li, L. Y. et al., A novel antiangiogenesis therapy using an integrin antagonist or anti-FLK-1 antibody<br />

coated Y-90-labeled nanoparticles, International Journal of Radiation Oncology Biology Physics,<br />

58, 1215–1227, 2004.<br />

31. Farokhzad, O. C. et al., Nanopartide-aptamer bioconjugates: A new approach for targeting prostate<br />

cancer cells, Cancer Research, 64, 7668–7672, 2004.<br />

32. Allen, T. M., Sapra, P., Moase, E., Moreira, J., and Iden, D., Adventures in targeting, Journal of<br />

Liposome Research, 12, 5–12, 2002.<br />

33. Christian, S. et al., Nucleolin expressed at the cell surface is a marker of endothelial cells in<br />

angiogenic blood vessels, Journal of Cell Biology, 163, 871–878, 2003.<br />

34. Porkka, K., Laakkonen, P., Hoffman, J. A., Bernasconi, M., and Ruoslahti, E., A fragment of the<br />

HMGN2 protein homes to the nuclei of tumor cells and tumor endothelial cells in vivo, Proceedings<br />

of the National Academy of Sciences of the United States of America, 99, 7444–7449, 2002.<br />

35. Pack, D. W., Hoffman, A. S., Pun, S., and Stayton, P. S., Design and development of polymers for<br />

gene delivery, Nature Reviews Drug Discovery, 4, 581–593, 2005.<br />

36. Frankel, A. D. and Pabo, C. O., Cellular uptake of the tat protein from human immunodeficiency<br />

virus, Cell, 55, 1189–1193, 1988.<br />

37. Wadia, J. S., Stan, R. V., and Dowdy, S. F., Transducible TAT-HA fusogenic peptide enhances escape of<br />

TAT-fusion proteins after lipid raft macropinocytosis, Nature Medicine, 10, 310–315, 2004.<br />

38. Kircher, M. F. et al., In vivo high resolution three-dimensional imaging of antigen-specific cytotoxic<br />

T-lymphocyte trafficking to tumors, Cancer Research, 63, 6838–6846, 2003.<br />

39. Weissleder, R., Bogdanov, A., Neuwelt, E. A., and Papisov, M., Long-circulating iron-oxides for<br />

Mr-imaging, Advanced Drug Delivery Reviews, 16, 321–334, 1995.<br />

40. Wagner, E., Application of membrane-active peptides for nonviral gene delivery, Advanced Drug<br />

Delivery Reviews, 38, 279–289, 1999.<br />

41. Plank, C., Oberhauser, B., Mechtler, K., Koch, C., and Wagner, E., The influence of endosomedisruptive<br />

peptides on gene-transfer using synthetic virus-like gene-transfer systems, Journal of<br />

Biological Chemistry, 269, 12918–12924, 1994.<br />

42. Mastrobattista, E., Crommelin, D. J. A., Wilschut, J., and Storm, G., Targeted liposomes for delivery of<br />

protein-based drugs into the cytoplasm of tumor cells, Journal of Liposome Research, 12, 57–65, 2002.<br />

43. Kakudo, T. et al., Transferrin-modified liposomes equipped with a pH-sensitive fusogenic peptide:<br />

An artificial viral-like delivery system, Biochemistry, 43, 5618–5628, 2004.<br />

44. Lanza, G. M. et al., Novel paramagnetic contrast agents for molecular imaging and targeted drug<br />

delivery, Current Pharmaceutical Biotechnology, 5, 495–507, 2004.<br />

45. West, J. L. and Halas, N. J., Engineered nanomaterials for biophotonics applications: Improving<br />

sensing, imaging, and therapeutics, Annual Review of Biomedical Engineering, 5, 285–292, 2003.<br />

46. Stark, D. D. et al., Superparamagnetic iron-oxide—clinical-application as a contrast agent for MR<br />

imaging of the liver, Radiology, 168, 297–301, 1988.<br />

47. Choi, H., Choi, S. R., Zhou, R., Kung, H. F., and Chen, I. W., Iron oxide nanoparticles as magnetic<br />

resonance contrast agent for tumor imaging via folate receptor-targeted delivery, Academic<br />

Radiology, 11, 996–1004, 2004.<br />

48. Huh, Y. M. et al., In vivo magnetic resonance detection of cancer <strong>by</strong> using multifunctional magnetic<br />

nanocrystals, Journal of the American Chemical Society, 127, 12387–12391, 2005.<br />

49. Wang, S. J., Brechbiel, M., and Wiener, E. C., Characteristics of a new MRI contrast agent<br />

prepared from polypropyleneimine dendrimers, generation 2, Investigative Radiology, 38,<br />

662–668, 2003.<br />

50. Lanza, G. M. et al., Molecular imaging and targeted drug delivery with a novel, ligand-directed<br />

paramagnetic nanoparticle technology, Academic Radiology, 9, S330–S331, 2002.<br />

51. Loo, C. et al., Nanoshell-enabled photonics-based imaging and therapy of cancer, Technology in<br />

Cancer Research & Treatment, 3, 33–40, 2004.<br />

52. Sokolov, K. et al., Real-time vital optical imaging of precancer using anti-epidermal growth factor<br />

receptor antibodies conjugated to gold nanoparticles, Cancer Research, 63, 1999–2004, 2003.<br />

53. Loo, C., Lowery, A., Halas, N., West, J., and Drezek, R., Immunotargeted nanoshells for integrated<br />

cancer imaging and therapy, Nano Letters, 5, 709–711, 2005.<br />

54. Parak, W. J. et al., Biological applications of colloidal nanocrystals, Nanotechnology, 14, R15–R27,<br />

2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Multifunctional Nanoparticles for Cancer Therapy 73<br />

55. Gao, X. H., Cui, Y. Y., Levenson, R. M., Chung, L. W. K., and Nie, S. M., In vivo cancer targeting<br />

and imaging with semiconductor quantum dots, Nature Biotechnology, 22, 969–976, 2004.<br />

56. Kim, S. et al., Near-infrared fluorescent type II quantum dots for sentinel lymph node mapping,<br />

Nature Biotechnology, 22, 93–97, 2004.<br />

57. Derfus, A. M., Chan, W. C. W., and Bhatia, S. N., Probing the cytotoxicity of semiconductor<br />

quantum dots, Nano Letters, 4, 11–18, 2004.<br />

58. Choi, Y. and Baker, J. R., Targeting cancer cells with DNA-assembled dendrimers—a mix and<br />

match strategy for cancer, Cell Cycle, 4, 669–671, 2005.<br />

59. Lu, Y., Yin, Y. D., Mayers, B. T., and Xia, Y. N., Modifying the surface properties of superparamagnetic<br />

iron oxide nanoparticles through a sol–gel approach, Nano Letters, 2, 183–186, 2002.<br />

60. Roy, I. et al., Ceramic-based nanoparticles entrapping water-insoluble photosensitizing anti-cancer<br />

drugs: A novel drug–carrier system for photodynamic therapy, Journal of the American Chemical<br />

Society, 125, 7860–7865, 2003.<br />

61. Lanza, G. M. et al., A novel site-targeted ultrasonic contrast agent with broad biomedical application,<br />

Circulation, 94, 3334–3340, 1996.<br />

62. Alkan-Onyuksel, H. et al., Development of inherently echogenic liposomes as an ultrasonic contrast<br />

agent, Journal of Pharmaceutical Sciences, 85, 486–490, 1996.<br />

63. Huang, S. L. et al., Improving ultrasound reflectivity and stability of echogenic liposomal dispersions<br />

for use as targeted ultrasound contrast agents, Journal of Pharmaceutical Sciences, 90, 1917–1926,<br />

2001.<br />

64. Rabin, O., Perez, J. M., Grimm, J., Wojtkiewicz, G., and Weissleder, R., An x-ray computed<br />

tomography imaging agent based on long-circulating bismuth sulphide nanoparticles, Nature<br />

Materials, 5, 118–122, <strong>2006</strong>.<br />

65. Kelly, K. A. et al., Detection of vascular adhesion molecule-1 expression using a novel multimodal<br />

nanoparticle, Circulation Research, 96, 327–336, 2005.<br />

66. Morawski, A. M., Lanza, G. A., and Wickline, S. A., Targeted contrast agents for magnetic resonance<br />

imaging and ultrasound, Current Opinion in Biotechnology, 16, 89–92, 2005.<br />

67. Perez, J. M., Josephson, L., O’Loughlin, T., Hogemann, D., and Weissleder, R., Magnetic relaxation<br />

switches capable of sensing molecular interactions, Nature Biotechnology, 20, 816–820, 2002.<br />

68. Perez, J. M., Simeone, F. J., Saeki, Y., Josephson, L., and Weissleder, R., Viral-induced selfassembly<br />

of magnetic nanoparticles allows the detection of viral particles in biological media,<br />

Journal of the American Chemical Society, 125, 10192–10193, 2003.<br />

69. Perez, J. M., Simeone, F. J., Tsourkas, A., Josephson, L., and Weissleder, R., Peroxidase substrate<br />

nanosensors for MR imaging, Nano Letters, 4, 119–122, 2004.<br />

70. Georganopoulou, D. G. et al., Nanoparticle-based detection in cerebral spinal fluid of a soluble<br />

pathogenic biomarker for Alzheimer’s disease, Proceedings of the National Academy of Sciences<br />

of the United States of America, 102, 2273–2276, 2005.<br />

71. Mirkin, C. A., Letsinger, R. L., Mucic, R. C., and Storhoff, J. J., A DNA-based method for rationally<br />

assembling nanoparticles into macroscopic materials, Nature, 382, 607–609, 1996.<br />

72. Stevens, M. M., Flynn, N. T., Wang, C., Tirrell, D. A., and Langer, R., Coiled-coil peptide-based<br />

assembly of gold nanoparticles, Advanced Materials, 16, 915–918, 2004.<br />

73. Harris, T. J., von Maltzahn, G., Derfus, A. M., Ruoslahti, E., and Bhatia, S. N., Proteolytic actuation<br />

of nanoparticle self-assembly, Angewandte Chemie-International Edition, In press.<br />

74. Muller, R. H. and Keck, C. M., Challenges and solutions for the delivery of biotech drugs—a review<br />

of drug nanocrystal technology and lipid nanoparticles, Journal of Biotechnology, 113, 151–170,<br />

2004.<br />

75. Duncan, R., The dawning era of polymer therapeutics, Nature Reviews Drug Discovery, 2, 347–360,<br />

2003.<br />

76. Sessa, G. and Weissman, G., Phospholipid spherules (liposomes) as a model for biological<br />

membranes, Journal of Lipid Research, 9(3): 310–318, 1968.<br />

77. Gabizon, A. and Martin, F., Polyethylene glycol coated (pegylated) liposomal doxorubicin—<br />

rationale for use in solid tumours, Drugs, 54, 15–21, 1997.<br />

78. Campbell, R. B., Balasubramanian, S. V., and Straubinger, R. M., Influence of cationic lipids on the<br />

stability and membrane properties of paclitaxel-containing liposomes, Journal of Pharmaceutical<br />

Sciences, 90, 1091–1105, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


74<br />

Nanotechnology for Cancer Therapy<br />

79. Sapra, P. and Allen, T. M., Improved outcome when B-cell lymphoma is treated with combinations<br />

of immunoliposomal anti-cancer drugs targeted to both the CD19 and CD20 epitopes, Clinical<br />

Cancer Research, 10, 4893, 2004.<br />

80. Torchilin, V. P. et al., p-nitrophenylcarbonyl-PEG-PE-liposomes: Fast and simple attachment of<br />

specific ligands, including monoclonal antibodies, to distal ends of PEG chains via p-nitrophenylcarbonyl<br />

groups, Biochimica et Biophysica Acta-Biomembranes, 1511, 397–411, 2001.<br />

81. Park, J. W. et al., Tumor targeting using anti-her2 immunoliposomes, Journal of Controlled Release,<br />

74, 95–113, 2001.<br />

82. Miyamoto, M. et al., Preparation of gadolinium-containing emulsions stabilized with phosphatidylcholine-surfactant<br />

mixtures for neutron-capture therapy, Chemical & Pharmaceutical Bulletin, 47,<br />

203–208, 1999.<br />

83. Panyam, J. and Labhasetwar, V., Biodegradable nanoparticles for drug and gene delivery to cells and<br />

tissue, Advanced Drug Delivery Reviews, 55, 329–347, 2003.<br />

84. Nakanishi, T. et al., Development of the polymer micelle carrier system for doxorubicin, Journal of<br />

Controlled Release, 74, 295–302, 2001.<br />

85. Yokoyama, M. et al., Selective delivery of adiramycin to a solid tumor using a polymeric micelle<br />

carrier system, Journal of Drug Targeting, 7, 171–186, 1999.<br />

86. Torchilin, V. P., Structure and design of polymeric surfactant-based drug delivery systems, Journal<br />

of Controlled Release, 73, 137–172, 2001.<br />

87. Discher, B. M. et al., Polymersomes: Tough vesicles made from diblock copolymers, Science, 284,<br />

1143–1146, 1999.<br />

88. Ahmed, F. and Discher, D. E., Self-porating polymersomes of PEG-PLA and PEG-PCL: Hydrolysistriggered<br />

controlled release vesicles, Journal of Controlled Release, 96, 37–53, 2004.<br />

89. Xu, J. P., Ji, J., Chen, W. D., and Shen, J. C., Novel biomimetic polymersomes as polymer therapeutics<br />

for drug delivery, Journal of Controlled Release, 107, 502–512, 2005.<br />

90. Patri, A. K., Majoros, I. J., and Baker, J. R., Dendritic polymer macromolecular carriers for drug<br />

delivery, Current Opinion in Chemical Biology, 6, 466–471, 2002.<br />

91. Kojima, C., Kono, K., Maruyama, K., and Takagishi, T., Synthesis of polyamidoamine dendrimers<br />

having poly(ethylene glycol) grafts and their ability to encapsulate anti-cancer drugs, Bioconjugate<br />

Chemistry, 11, 910–917, 2000.<br />

92. Patri, A. K. et al., Synthesis and in vitro testing of J591 antibody-dendrimer conjugates for targeted<br />

prostate cancer therapy, Bioconjugate Chemistry, 15, 1174–1181, 2004.<br />

93. Tripathi, P. K. et al., Dendrimer grafts for delivery of 5-flurouracil, Pharmazie, 57, 261–264, 2002.<br />

94. Quintana, A. et al., Design and function of a dendrimer-based therapeutic nanodevice targeted to<br />

tumor cells through the folate receptor, Pharmaceutical Research, 19, 1310–1316, 2002.<br />

95. Vinogradov, S. V., Batrakova, E. V., and Kabanov, A. V., Nanogels for oligonucleotide delivery to<br />

the brain, Bioconjugate Chemistry, 15, 50–60, 2004.<br />

96. Jordan, A., Scholz, R., Wust, P., Fahling, H., and Felix, R., Magnetic fluid hyperthermia (MFH):<br />

Cancer treatment with AC magnetic field induced excitation of biocompatible superparamagnetic<br />

nanoparticles, Journal of Magnetism and Magnetic Materials, 201, 413–419, 1999.<br />

97. Pankhurst, Q. A., Connolly, J., Jones, S. K., and Dobson, J., Applications of magnetic nanoparticles<br />

in biomedicine, Journal of Physics D—Applied Physics, 36, R167–R181, 2003.<br />

98. Derfus, A. M. and Bhatia, S. N., Unpublished data.<br />

99. Rabin, Y., Is intracellular hyperthermia superior to extracellular hyperthermia in the thermal sense?<br />

International Journal of Hyperthermia, 18, 194–202, 2002.<br />

100. Hirsch, L. R. et al., Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic<br />

resonance guidance, Proceedings of the National Academy of Sciences of the United States of<br />

America, 100, 13549–13554, 2003.<br />

101. O’Neal, D. P., Hirsch, L. R., Halas, N. J., Payne, J. D., and West, J. L., Photo-thermal tumor ablation<br />

in mice using near infrared-absorbing nanoparticles, Cancer Letters, 209, 171–176, 2004.<br />

102. Plank, C. et al., The magnetofection method: Using magnetic force to enhance gene delivery,<br />

Biological Chemistry, 384, 737–747, 2003.<br />

103. Nishiyama, N. et al., Light-harvesting ionic dendrimer porphyrins as new photosensitizers for<br />

photodynamic therapy, Bioconjugate Chemistry, 14, 58–66, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Multifunctional Nanoparticles for Cancer Therapy 75<br />

104. Konan, Y. N., Gurny, R., and Allemann, E., State of the art in the delivery of photosensitizers for<br />

photodynamic therapy, Journal of Photochemistry and Photobiology B—Biology, 66, 89–106, 2002.<br />

105. Kopelman, R. et al., Multifunctional nanoparticle platforms for in vivo MRI enhancement and<br />

photodynamic therapy of a rat brain cancer, Journal of Magnetism and Magnetic Materials, 293,<br />

404–410, 2005.<br />

106. Crowder, K. C. et al., Sonic activation of molecularly-targeted nanoparticles accelerates transmembrane<br />

lipid delivery to cancer cells through contact-mediated mechanisms: Implications for<br />

enhanced local drug delivery, Ultrasound in Medicine and Biology, 31, 1693–1700, 2005.<br />

107. Gao, Z. G., Fain, H. D., and Rapoport, N., Controlled and targeted tumor chemotherapy <strong>by</strong> micellarencapsulated<br />

drug and ultrasound, Journal of Controlled Release, 102, 203–222, 2005.<br />

108. Hainfeld, J. F., Slatkin, D. N., and Smilowitz, H. M., The use of gold nanoparticles to enhance<br />

radiotherapy in mice, Physics in Medicine and Biology, 49, N309–N315, 2004.<br />

109. Haag, R., Supramolecular drug-delivery systems based on polymeric core-shell architectures, Angewandte<br />

Chemie-International Edition, 43, 278–282, 2004.<br />

110. Murthy, N., Campbell, J., Fausto, N., Hoffman, A. S., and Stayton, P. S., Design and synthesis of<br />

pH-responsive polymeric carriers that target uptake and enhance the intracellular delivery of oligonucleotides,<br />

Journal of Controlled Release, 89, 365–374, 2003.<br />

111. Forster, S., Abetz, V., and Muller, A. H. E., Polyelectrolyte block copolymer micelles, Polyelectrolytes<br />

with Defined Molecular Architecture Ii, 166, 173–210, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


6<br />

CONTENTS<br />

Neutron Capture Therapy of<br />

Cancer: Nanoparticles and High<br />

Molecular Weight Boron<br />

Delivery Agents<br />

Gong Wu, Rolf F. Barth, Weilian Yang, Robert J. Lee,<br />

Werner Tjarks, Marina V. Backer, and Joseph M. Backer<br />

6.1 Introduction ......................................................................................................................... 78<br />

6.2 General Requirements for Boron Delivery Agents ............................................................ 78<br />

6.3 Low-Molecular-Weight Delivery Agents ........................................................................... 79<br />

6.4 High-Molecular-Weight Boron Delivery Agents ............................................................... 79<br />

6.5 Dendrimer-Related Delivery Agents .................................................................................. 80<br />

6.5.1 Properties of Dendrimers ........................................................................................ 80<br />

6.5.2 Boronated Dendrimers Linked to Monoclonal Antibodies .................................... 80<br />

6.5.2.1 Boron Clusters Directly Linked to mAb ................................................. 80<br />

6.5.2.2 Attachment of Boronated Dendrimers to mAb ....................................... 81<br />

6.5.3 Boronated Dendrimers Delivered <strong>by</strong> Receptor Ligands ........................................ 81<br />

6.5.3.1 Epidermal Growth Factors (EGF) ........................................................... 81<br />

6.5.3.2 Folate Receptor-Targeting Agents........................................................... 83<br />

6.5.3.3 Vascular Endothelial Growth Factor (VEGF) ......................................... 83<br />

6.5.4 Other Boronated Dendrimers .................................................................................. 85<br />

6.6 Liposomes as Boron Delivery Agents ................................................................................ 86<br />

6.6.1 Overview of Liposomes .......................................................................................... 86<br />

6.6.2 Boron Delivery <strong>by</strong> Nontargeted Liposomes ........................................................... 86<br />

6.6.2.1 Liposomal Encapsulation of Sodium Borocaptate and<br />

Boronophenylalanine ............................................................................... 86<br />

6.6.2.2 Liposomal Encapsulation of Other Boranes and Carboranes ................. 87<br />

6.6.3 Boron Delivery <strong>by</strong> Targeted Liposomes ................................................................ 89<br />

6.6.3.1 Immunoliposomes .................................................................................... 89<br />

6.6.3.2 Folate Receptor-Targeted Liposomes ...................................................... 90<br />

6.6.3.3 EGFR-Targeted Liposomes ..................................................................... 91<br />

6.7 Boron Delivery <strong>by</strong> Dextrans ............................................................................................... 92<br />

Experimental studies described in this report were supported <strong>by</strong> N.I.H. grants 1 R01 CA098945 (R.F. Barth), 1 R01<br />

CA79758 (W. Tjarks), and Department of Energy grants DE-FG02-98ER62595 (R.F. Barth) and DE-FG-2-02FR83520<br />

(J.M. Backer).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

77


78<br />

6.8 Other Macromolecules Used for Delivering Boron Compounds....................................... 93<br />

6.9 Delivery of Boron-Containing Macromolecules to Brain Tumors .................................... 93<br />

6.9.1 General Considerations ........................................................................................... 93<br />

6.9.2 Drug-Transport Vectors .......................................................................................... 93<br />

6.9.3 Direct Intracerebral Delivery .................................................................................. 94<br />

6.9.4 Convection-Enhanced Delivery .............................................................................. 94<br />

6.10 Clinical Considerations and Conclusions ......................................................................... 95<br />

6.11 Summary ........................................................................................................................... 96<br />

Acknowledgments ........................................................................................................................ 96<br />

References .................................................................................................................................... 96<br />

6.1 INTRODUCTION<br />

After decades of intensive research, high-grade gliomas, and specifically glioblastoma multiforme<br />

(GBM), are still extremely resistant to all current forms of therapy including surgery,<br />

chemotherapy, radiotherapy, immunotherapy and gene therapy. 1–5 The five-year survival rate of<br />

patients diagnosed with GBM in the United States is less than a few percent 6,7 despite aggressive<br />

treatment using combinations of therapeutic modalities. This is due to the infiltration of malignant<br />

cells beyond the margins of resection and their spread into both gray and white matter <strong>by</strong> the time of<br />

surgical resection. 8,9 High-grade gliomas are histologically complex and heterogeneous in their<br />

cellular composition. Recent molecular genetic studies of gliomas have shown how complex the<br />

development of these tumors is. 10 Glioma cells and their neoplastic precursors have biologic<br />

properties that allow them to evade a tumor associated host immune response, 11 and biochemical<br />

properties that allow them to invade the unique extracellular environment of the brain. 12,13 Consequently,<br />

high-grade supratentorial gliomas must be regarded as a whole-brain disease. 14 The<br />

inability of chemo- and radiotherapy to cure patients with high-grade gliomas is due to their<br />

failure to eradicate micro-invasive tumor cells within the brain. To successfully treat these<br />

tumors, strategies must be developed that can selectively target malignant cells with little or no<br />

effect on normal cells and tissues adjacent to the tumor. Boron neutron capture therapy (BNCT) is<br />

based on the nuclear capture and fission reactions that occur when non radioactive boron-10 is<br />

irradiated with low-energy thermal neutrons to yield high-linear-energy-transfer (LET) alpha<br />

particles ( 4 He) and recoiling lithium-7 ( 7 Li) nuclei. For BNCT to be successful, a sufficient<br />

number of 10 B atoms (approximately 10 9 atoms/cell) must be delivered selectively to the tumor<br />

and enough thermal neutrons must be absorbed <strong>by</strong> them to sustain a lethal 10 B(n,a) 7 Li-capture<br />

reaction. The destructive effects of these high-energy particles are limited to boron-containing cells.<br />

BNCT primarily has been used to treat high-grade gliomas, 15,16 and either cutaneous primaries 17 or<br />

cerebral metastases of melanoma. 18 More recently, it also has been used to treat patients with headand-neck<br />

19,20 and metastatic liver cancer. 21,22 BNCT is a biologically rather than physically<br />

targeted type of radiation treatment. If sufficient amounts of 10 B and thermal neutrons can be<br />

delivered to the target volume, the potential exists to destroy tumor cells dispersed in the normal<br />

tissue parenchyma. Readers interested in more in depth coverage of other topics related to<br />

BNCT are referred to several recent reviews 23–25 and monographs. 15,24 This review will focus<br />

on boron-containing macromolecules and nanovehicles as boron delivery agents.<br />

6.2 GENERAL REQUIREMENTS FOR BORON DELIVERY AGENTS<br />

Nanotechnology for Cancer Therapy<br />

A successful boron delivery agent should have (1) no or minimal systemic toxicity with rapid<br />

clearance from blood and normal tissues, (2) high tumor (approximately 20 mg 10 B/g) and low<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 79<br />

normal tissue uptake, (3) high tumor to brain (T:Br) and tumor to blood (T:Bl) concentration ratios<br />

(greater than 3–4:1), and (4) persistence in the tumor for a sufficient period of time to carry out<br />

BNCT. At this time, no single boron delivery agent fulfills all of these criteria. However, as a result<br />

of new synthetic techniques and increased knowledge of the biological and biochemical requirements<br />

for an effective agent, multiple new boron agents have emerged and these are described in a<br />

special issue of Anti-Cancer Agents in Medical Chemistry (6, 2, <strong>2006</strong>). The major challenge in their<br />

development has been the requirement for specific tumor targeting to achieve boron concentrations<br />

sufficient to deliver therapeutic doses of radiation to the tumor with minimal normal-tissue toxicity.<br />

The selective destruction of GBM cells in the presence of normal cells represents an even greater<br />

challenge compared to malignancies at other anatomic sites.<br />

6.3 LOW-MOLECULAR-WEIGHT DELIVERY AGENTS<br />

In the 1950s and early 1960s clinical trials of BNCT were carried out using boric acid and some of<br />

its derivatives as delivery agents. These simple chemical compounds had poor tumor retention,<br />

attained low T:Br ratios and were nonselective. 26,27 Among the hundreds of low-molecular-weight<br />

(LMW) boron-containing compounds that were synthesized, two appeared to be promising. One,<br />

based on arylboronic acids, 28 was (L)-4-dihydroxy-borylphenylalanine, referred to as boronophenylalanine<br />

or BPA (Figure 6.1, 1). The second, a polyhedral borane anion, was sodium<br />

mercaptoundecahydro-closo-dodecaborate, 29 more commonly known as sodium borocaptate or<br />

BSH (2). These two compounds persisted longer in animal tumors compared with related<br />

molecules, attained T:Br and T:Bl boron ratios greater than 1 and had low toxicity. 10 B-enriched<br />

BPA, complexed with fructose to improve its water solubility, and BSH, have been used clinically<br />

for BNCT of brain, as well as extracranial tumors. Although their selective accumulation in tumors<br />

is not ideal, the safety of these two drugs following intravenous (i.v.) administration has been well<br />

established. 30,31<br />

6.4 HIGH-MOLECULAR-WEIGHT BORON DELIVERY AGENTS<br />

High-molecular-weight (HMW) delivery agents usually contain a stable boron group or cluster<br />

linked via a hydrolytically stable bond to a tumor-targeting moiety, such as monoclonal antibodies<br />

(mAbs) or LMW receptor-targeting ligands. Examples of these include epidermal growth factor<br />

(HO) 2 B<br />

OCN<br />

1<br />

3<br />

H<br />

CO 2 –<br />

NH 3 +<br />

N(CH 3) 3Na<br />

FIGURE 6.1 Structures of two compounds used clinically for boron neutron capture therapy (BNCT), 4-dihydroxyborylphenylalanine<br />

or boronophenylalanine (BPA, 1), and disodium undecahydro-closo-dodecaborate or<br />

sodium borocaptate (BSH, 2) and the isocyanato polyhedral borane (3), which has been used to heavily<br />

boronate dendrimers.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Na 2<br />

SH<br />

2<br />

BH<br />

B


80<br />

(EGF) and the mAb cetuximab (IMC-C225) to target the EGF receptor (EGFR) or its mutant<br />

isoform EGFRvIII, which are overexpressed in a variety of malignant tumors including gliomas,<br />

and squamous-cell carcinomas of the head and neck. 32 Agents that are to be administered systemically<br />

should be water soluble (WS), but lipophilicity is important to cross the blood–brain barrier<br />

(BBB) and diffuse within the brain and the tumor. There should be a favorable differential in boron<br />

concentrations between tumor and normal brain, there<strong>by</strong> enhancing their tumor specificity. Their<br />

amphiphilic character is not as crucial for LMW agents that target-specific biological transport<br />

systems and/or are incorporated into nanovehicles such as liposomes. Molecular weight also is an<br />

important factor because it determines the rate of diffusion both within the brain and the tumor.<br />

Detailed reviews of the state-of-the-art of compound development for BNCT have been<br />

published. 33,34 In this review, we will focus on boron containing macromolecules and liposomes<br />

as delivery agents for BNCT, and how they can be most effectively administered.<br />

6.5 DENDRIMER-RELATED DELIVERY AGENTS<br />

6.5.1 PROPERTIES OF DENDRIMERS<br />

Dendrimers are synthetic polymers with a well-defined globular structure. As shown in Figure 6.2,<br />

they are composed of a core molecule, repeat units that have three or more functionalities, and<br />

reactive surface groups. 35,36 Two techniques have been used to synthesize these macromolecules:<br />

divergent growth outwards from the core, 37 or convergent growth from the terminal groups inwards<br />

towards the core. 36,38 Regular and repeated branching at each monomer group gives rise to a<br />

symmetric structure and pattern to the entire globular dendrimers. Dendrimers are an attractive<br />

platform for macromolecular imaging and gene delivery because of their low cytotoxicity and their<br />

multiple types of reactive terminal groups. 36,39–44<br />

6.5.2 BORONATED DENDRIMERS LINKED TO MONOCLONAL ANTIBODIES<br />

6.5.2.1 Boron Clusters Directly Linked to mAb<br />

Monoclonal antibodies (mAb) have been attractive targeting agents for delivering radionuclides, 45<br />

drugs, 46–50 toxins, 51 and boron to tumors. 52–55 Prior to the introduction of dendrimers as boron<br />

carriers, boron compounds were directly attached to mAbs. 53,54 It has been calculated that<br />

Terminal groups<br />

R HN<br />

Linkers<br />

= Repeating units<br />

R = mAb, EGF, Folic acid or VEGF<br />

H<br />

N H<br />

N –<br />

N(CH3 ) 3<br />

FIGURE 6.2 Structure of a boronated PAMAM dendrimer that has been linked to targeting moieties.<br />

PAMAM dendrimers consist of a core, repeating polyamido amino units, and reactive terminal groups.<br />

Each successively higher generation of PAMAM dendrimer has a geometrically incremental number of<br />

terminal groups. Boronated dendrimers have been prepared <strong>by</strong> reacting them with water-soluble isocyanato<br />

polyhedral boranes and subsequently attaching them to targeting moieties <strong>by</strong> means of heterobifunctional<br />

linkers.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

O<br />

BH<br />

B


Neutron Capture Therapy of Cancer 81<br />

approximately 10 9 10 B atoms per cell (approximately 20 mg/g tumor) must be delivered to kill<br />

tumor cells. 55,56 Based on the assumption of 10 6 antigenic receptor sites per cell, approximately 50–<br />

100 boron cage structures of carboranes, or polyhedral borane anions and their derivatives must be<br />

linked to each mAb molecule to deliver the required amount of boron for NCT. The attachment of<br />

such a large number of boron cages to a mAb may result in precipitation of the bioconjugate or a<br />

loss of its immunological activity. Solubility can be improved <strong>by</strong> inserting a water-soluble gluconamide<br />

group into the protein-binding boron cage compounds, there<strong>by</strong> enhancing their water<br />

solubility. 57 This modification makes it possible to incorporate up to 1100 boron atoms into a<br />

human gamma globulin (HGG) molecule without any precipitation. Other approaches to<br />

enhance solubility include the use of negatively charged carboranes 58 or polyhedral borane<br />

anions, 59 as well as the insertion of carbohydrate groups. 60,61 A major limitation of using an<br />

agent containing a single boron cage is that a large number of sites must be modified to deliver<br />

10 3 boron atoms per molecule of antibody and this can reduce its immunoreactivity activity. Alam<br />

et al. showed that attachment of an average of 1300 boron atoms to mAb 17-1A, which is directed<br />

against human colorectal carcinoma cells, resulted in a 90% loss of its immunoreactivity. 62<br />

6.5.2.2 Attachment of Boronated Dendrimers to mAb<br />

Dendrimers are one of the most attractive polymers that have been used as boron carriers due to<br />

their well-defined structure and large number of reactive terminal groups. Depending on the antigen<br />

site density, approximately 1000 boron atoms need to be attached per molecule of dendrimer and<br />

subsequently linked to the mAb. In our first study, second- and fourth-generation polyamido amino<br />

(PAMAM or “starburst”) dendrimers, which have 12 and 48 reactive terminal amino groups,<br />

respectively, were reacted with the water-soluble isocyanato polyhedral borane [Na(CH3)3NB10<br />

H8NCO] (3). 63,64 The boronated dendrimer then was linked to the mAb IB16-6, which is directed<br />

against the murine B16 melanoma, <strong>by</strong> means of two heterobifunctional linkers, m-maleimidobenzoyl-N-hydroxysulfosuccinimide<br />

ester (sulfo-MBS) and N-succinimidyl 3-(2-pyridyldithio)<br />

propionate (SPDP). 63,65 However, following i.v. administration, large amounts of the bioconjugate<br />

accumulated in the liver, and spleen and it was concluded that random conjugation of boronated<br />

dendrimers to a mAb could alter its binding affinity and biodistribution. To minimize the loss of<br />

mAb reactivity, a fifth-generation PAMAM dendrimer was boronated with the same polyhedral<br />

borane anion, and more recently it was site-specifically linked to the anti-EGFR mAb cutuximab<br />

(or IMC-C225) or the EGFRvIII-specific mAb L8A4 (Figure 6.3). Cetuximab was linked via<br />

glycosidic moieties in the Fc region <strong>by</strong> means of two heterobifunctional reagents, SPDP and<br />

N-(k-maleimidoundecanoic acid) hydrazide (KMUH). 66,67 The resulting bioconjugate, designated<br />

C225-G5-B 1100, contained approximately 1100 boron atoms per molecule of cetuximab and<br />

retained its aqueous solubility in 10% DMSO and its in vitro and in vivo immunoreactivity. As<br />

determined <strong>by</strong> a competitive binding assay, there was a less than 1 log unit decrease in affinity for<br />

EGFR (C) glioma cell line F98EGFR, compared to that of unmodified cetuximab. 66 In vivo biodistribution<br />

studies, carried out 24 h after intratumoral (i.t) administration of the bioconjugate,<br />

demonstrated that 92.3 mg/g of boron was retained in rats bearing F98EGFR gliomas, compared to<br />

36.5 mg/g in EGFR (K) F98 parental tumors and 6.7 mg/g in normal brain. 67<br />

6.5.3 BORONATED DENDRIMERS DELIVERED BY RECEPTOR LIGANDS<br />

6.5.3.1 Epidermal Growth Factors (EGF)<br />

Due to its increased expression in a variety of tumors, including high-grade gliomas, and its low or<br />

undetectable expression in normal brain, EGFR is an attractive target for cancer therapy. 68–70<br />

As described above, targeting of EGFR has been carried out using either mAbs or alternatively,<br />

as described in this section, EGF, which is a single-chain, 53-mer heat and acid stable polypeptide.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


82<br />

HO<br />

O<br />

HO<br />

O<br />

O<br />

O<br />

O O<br />

O<br />

H<br />

N S<br />

O<br />

H<br />

N H N<br />

O<br />

O<br />

OH<br />

OH<br />

BH<br />

B<br />

Cetuximab (C225)<br />

NalO 4<br />

C225-CHO<br />

N(CH 3) 3 –<br />

KMUH-G5-B 1100<br />

O O<br />

O O<br />

N<br />

O<br />

(CH2 ) 10 N<br />

H<br />

N<br />

O<br />

O OH<br />

C225-G5-B 1100<br />

NH 2<br />

NH 2<br />

5th generation PAMAM dendrimer<br />

O<br />

S<br />

S<br />

NH 2<br />

H<br />

N<br />

H<br />

N<br />

N<br />

O<br />

SPDP-G5<br />

H<br />

N SH<br />

O<br />

HS-G5-B1100 O<br />

H<br />

N<br />

O<br />

N<br />

O<br />

O<br />

NH2 (CH2 ) 10 N<br />

H<br />

S<br />

KMUH-G5-B 1100<br />

S S N<br />

Na(CH 3 ) 3 NB 10 H 8 NCO<br />

O<br />

SPDP-G5-B 1100<br />

DTT<br />

S S N<br />

O<br />

N<br />

O<br />

O<br />

N<br />

O<br />

(CH 2 ) 10<br />

FIGURE 6.3 Conjugation scheme for linkage of a boron-containing dendrimer to the monoclonal antibody,<br />

C225 (cetuximab), which is directed against EGFR. (From Wu, G. et al., Bioconjug. Chem., 15, 185, 2004;<br />

Barth, R. F. et al., Appl. Radiat. Isot. 61, 899, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

O<br />

N<br />

H<br />

NH 2


Neutron Capture Therapy of Cancer 83<br />

It binds to a transmembrane glycoprotein with tyrosine kinase activity, which triggers dimerization<br />

and internalization. 71,72 Because the EGF boron bioconjugates have a much smaller MW than mAb<br />

conjugates, they should be capable of more rapid and effective tumor targeting than has been<br />

observed with mAbs. 67,73<br />

The procedure used to conjugate EGF to a boronated dendrimer was slightly different from that<br />

used to boronate mAbs. A fourth-generation PAMAM dendrimer was reacted with the isocyanato<br />

polyhedral borane Na(CH3)3NB10H8NCO. Next, reactive thiol groups were introduced into the<br />

boronated dendrimer using SPDP, and EGF was derivatized with sulfo-MBS. The reaction of thiol<br />

groups of the derivatized, boronated starburst dendrimer (BSD) with maleimide groups produced a<br />

stable BSD–EGF bioconjugate, which contained approximately 960 atoms of boron per molecule of<br />

EGF. 74 The BSD–EGF initially was bound to the cell surface membrane and then was endocytosed,<br />

which resulted in accumulation of boron in lysosomes. 74 Subsequently, in vitro and in vivo studies<br />

were carried out to evaluate the potential efficacy of the bioconjugate as a boron delivery agent for<br />

BNCT. 73 As will be described in more detail later on in this review, therapy studies demonstrated<br />

that F98EGFR-glioma-bearing rats that received either boronated EGF or mAb <strong>by</strong> either direct i.t.<br />

injection or convection enhanced delivery into the brain had a longer mean survival time (MST)<br />

than animals bearing F98 parental tumors following BNCT. 75–77<br />

6.5.3.2 Folate Receptor-Targeting Agents<br />

Folate receptor (FR) is overexpressed on a variety of human cancers, including those originating in<br />

ovary, lung, breast, endometrium and kidney. 78–80 Folic acid (FA) is a vitamin that is transported<br />

into cells via FR mediated endocytosis. It has been well documented that the attachment of FA via<br />

its g-carboxylic function to other molecules does not alter its endocytosis <strong>by</strong> FR-expressing cells. 81<br />

FR targeting has been used successfully to deliver protein toxins, chemotherapeutic, radioimaging,<br />

therapeutic and MRI contrast agents, 82 liposomes, 83 gene transfer vectors, 84 antisense oligonucleotides,<br />

85 ribozymes, and immunotherapeutic agents to FR-positive cancers. 86 To deliver boron<br />

compounds, FA was conjugated to heavily boronated third generation PAMAM dendrimers<br />

containing polyethylene glycol (PEG). 87 PEG was introduced into the bioconjugate to reduce its<br />

uptake <strong>by</strong> the reticuloendothelial system (RES), and more specifically, the liver and spleen. It was<br />

observed that folate linked to third generation PAMAM dendrimers containing 12–15 decaborate<br />

clusters and 1–1.5 PEG2000 units had the lowest hepatic uptake in C57Bl/6 mice (7.2–7.7% injected<br />

dose [I.D.]/g liver). In vitro studies using FR (C) KB cells demonstrated receptor-dependent uptake<br />

of the bioconjugate. Biodistribution studies with this conjugate, carried out in C57Bl/6 mice<br />

bearing subcutaneous (s.c.) implants of the FR (C) murine sarcoma 24JK-FBP, demonstrated<br />

selective tumor uptake (6.0% I.D./g tumor), but there was high hepatic (38.8% I.D./g) and renal<br />

(62.8% I.D./g) uptake. 87<br />

6.5.3.3 Vascular Endothelial Growth Factor (VEGF)<br />

There is preclinical and clinical evidence indicating that angiogenesis plays a major role in the<br />

growth and dissemination of malignant tumors. 88,89 Inhibition of angiogenesis has yielded<br />

promising results in a number of experimental animal tumor models. 90,91 The most important<br />

molecular targets have been vascular endothelial growth factor (VEGF) and its receptor<br />

(VEGFR). 92,93 We recently have constructed a human VEGF121 isoform fused to a novel 15-aa<br />

cysteine-containing tag designed for site-specific conjugation of therapeutic and diagnostic<br />

agents 94 (Figure 6.4). A boronated fifth-generation PAMAM dendrimer (BD), tagged with a<br />

near-infrared (IR) fluorescent dye Cy5, was conjugated using the heterobifunctional reagent<br />

sulfo-LC–SPDP to BD/Cy5 via reactive SH-groups generated in the VEGF fusion protein <strong>by</strong><br />

mild dithiothreitol (DTT) treatment. The bioconjugate, designated VEGF–BD/Cy5, contained<br />

1050–1100 boron atoms per dimeric VEGF molecule. VEGF–BD/Cy5 retained the in vitro<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


84<br />

H 2N<br />

Cy5<br />

Cy5<br />

A<br />

B<br />

G5<br />

G5<br />

G5<br />

Cy5<br />

NH 2<br />

NH 2<br />

G5<br />

NH 2<br />

SPDP<br />

SPDP<br />

Cy5<br />

S<br />

S<br />

G5<br />

SH<br />

SH<br />

D<br />

S<br />

S<br />

VEGF-BD-Cy5<br />

H<br />

N H N<br />

functional activity of VEGF121, which was similar to that of native VEGF. Tagging the bioconjugate<br />

with Cy5 dye permitted near-IR fluorescence imaging of its in vitro and in vivo uptake.<br />

In vitro uptake was determined <strong>by</strong> incubating VEGF–BD/Cy5 with VEGFR-2 overexpressing<br />

PAE/KDR cells and in vivo uptake was evaluated in mice bearing s.c. tumor implants. In vitro,<br />

HS<br />

G5<br />

S<br />

S<br />

O<br />

SH<br />

Cy5<br />

G5<br />

C<br />

Nanotechnology for Cancer Therapy<br />

BH<br />

B<br />

Cy5<br />

SH<br />

N(CH 3 ) 3 –<br />

FIGURE 6.4 Preparation of a vascular endothelial growth factor (VEGF) receptor-targeting nanovehicle<br />

(VEGF–BD/Cy5) (A). A fifth-generation (G5) PAMAM dendrimer initially is reacted with SPDP and the<br />

near-IR dye, Cy5-NHS, (A) dissolved in dimethylformamide/methanol mixture for 1 h. Following this, it is<br />

reacted with the polyisocynato polyhedral borane, Na(CH 3) 3–NB 10H 8NCO, in a carbonate buffer (pH 9.0)/9%<br />

acetone mixture for 24 h (B). Then a 1.5-fold molar excess of DTT is added to the VEGF monomer in a<br />

solution containing 20 mmol/L NaOAc (pH 6.5), 0.5 mol/L urea and 0.5 mol/L NDSB-221 and incubated at<br />

48C for 16 h (C). Finally, the boronated dendrimer is incubated with the targeting vehicle for 15 min to produce<br />

the VEGF–BD/Cy5 (D). (From Backer, M. V. et al., Mol. Cancer Ther., 4, 1423, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

SH


Neutron Capture Therapy of Cancer 85<br />

the bioconjugate localized in the perinuclear region and in vivo it primarily localized in regions of<br />

active angiogenesis. Furthermore, depletion of VEGFR-2 overexpressing cells from the tumor<br />

vasculature with VEGF-toxin fusion protein significantly decreased bioconjugate uptake, indicating<br />

that these cells were the primary targets of VEGF–BD/Cy5. These studies demonstrated that<br />

VEGF–BD/Cy5 potentially could be used as a diagnostic agent. 94 Further studies are planned to<br />

evaluate its therapeutic efficacy using the F98 rat glioma model, which we have used extensively<br />

to evaluate both LMW and HMW boron delivery agents. 95<br />

6.5.4 OTHER BORONATED DENDRIMERS<br />

An alternative method to deliver boron compounds <strong>by</strong> means of dendrimers is to incorporate<br />

carborane cages within the dendrimer (Figure 6.5). Matthew et al. have reported that 4, 8, or 16<br />

carboranes could be inserted into an aliphatic polyester dendrimer <strong>by</strong> means of a highly effective<br />

HO<br />

HO<br />

HO<br />

HO<br />

HO<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

OH<br />

O<br />

O<br />

O O<br />

O O<br />

O<br />

O O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

OH<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O O O<br />

FIGURE 6.5 Structure of a barborane-containing aliphatic polyester dendrimer. Carborane cages were incorporated<br />

into the interior of the dendrimer structure and the peripheral hydrophilic groups improved water<br />

solubility and were available for modification.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

HO<br />

O<br />

O O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

OH<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

OH<br />

OH<br />

BH<br />

C<br />

OH<br />

OH<br />

OH


86<br />

synthon, a bifunctional carborane derivative with an acid group and a benzyl ether protected<br />

alcohol. 96 The procedure employed a divergent synthesis with high yield. The resulting polyhydroxylated<br />

dendrimer was WS with a minimum ratio of eight hydroxyl groups per carborane cage.<br />

Carborane containing dendrimers potentially could be used as boron delivery agents for BNCT<br />

because it is possible to control the number of carborane moieties and overall solubility.<br />

6.6 LIPOSOMES AS BORON DELIVERY AGENTS<br />

6.6.1 OVERVIEW OF LIPOSOMES<br />

Liposomes are biodegradable, nontoxic vesicles that have been used to deliver both hydrophilic and<br />

hydrophobic agents (Figure 6.6). 97 Both classical and PEGylated (“stealth”) liposomes can increase<br />

the amounts of anti-cancer drugs that can be delivered to solid tumors <strong>by</strong> passive targeting. Rapidly<br />

growing solid tumors have increased permeability to nanoparticles due to increased capillary pore<br />

size. These can range from 100 to 800 nm. In comparison, endothelial pore size of normal tissues,<br />

which are impermeable to liposomes, can range in size from 60 to 80 nm. In addition, tumors lack<br />

efficient lymphatic drainage, and consequently, clearance of extravasated liposomes is slow. 98<br />

Modification of the liposomal surface <strong>by</strong> PEGylation or attachment of antibodies or receptor<br />

ligands, will improve their selective targeting and increase their circulation time. 98<br />

6.6.2 BORON DELIVERY BY NONTARGETED LIPOSOMES<br />

6.6.2.1 Liposomal Encapsulation of Sodium Borocaptate and Boronophenylalanine<br />

Liposomes have been extensively evaluated as nanovehicles for the delivery of boron compounds<br />

for NCT. 99,100 In vitro and in vivo studies have demonstrated that they can effectively and selectively<br />

deliver large quantities of boron to tumors and that the compounds delivered <strong>by</strong> liposomes<br />

have a longer tumor retention time. BPA is an amino acid analogue that is preferentially taken up <strong>by</strong><br />

cells with increased metabolic activity, such as tumor cells of varying histopathologic types<br />

Bilayer memberane Internal aqueous space<br />

Nanotechnology for Cancer Therapy<br />

5nm<br />

Hydrocarbon tails<br />

Polar head groups<br />

mAb or ligand<br />

FIGURE 6.6 Schematic diagram of the structure of a liposome that has an aqueous core and a lipid bilayer<br />

membrane. The latter is composed of polar head groups with hydrocarbon tails. The liposomal surface can be<br />

modified <strong>by</strong> PEGylation to prolong its circulation time and linked to either a mAb or a ligand for targeting.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

PEG


Neutron Capture Therapy of Cancer 87<br />

including melanomas, 31,101 gliomas 102 , and squamous cell carcinomas. 103,104 Because of its low<br />

aqueous solubility, BPA has been used as a fructose complex, which has permitted it to be administered<br />

i.v. rather than orally. 105,106 Following i.v. administration of BPA, which had been<br />

incorporated into conventional liposomes, there was rapid elimination <strong>by</strong> the RES with very low<br />

blood boron concentrations at 3 h. In contrast, if BPA was incorporated into liposomes composed of<br />

distearoyl phosphoethanolamine (DSPE)–PEG, therapeutically effective tumor boron concentrations<br />

(greater than 20 mg/g) were seen at 3 h and at 6 h, indicating that PEG-liposomes had<br />

evaded the RES. 107 In addition, BPA has been incorporated into the lipid bilayer of liposomes,<br />

composed of positively charged lipid 1,2-dioleoyl-3-trimethylammonium-propane (DOTAP) and<br />

the zwitterionic lipid, 1,2-dioleoyl-sn-glycerol-3-phosphoethanolamine (DOPE). 108 Cationic liposomes<br />

have been widely used as carriers of biomolecules that specifically target the cell nucleus, 109<br />

which would be advantageous for BNCT. Another clinically used drug, BSH, has been incorporated<br />

into liposomes composed of dipalmitoylphosphatidylcholine (DPPC)/Chol in a 1:1 molar ratio with<br />

and without PEG stabilization. 110 The average diameter of liposomes containing BSH was in the<br />

range of 100–110 nm. Both types of liposomes resulted in a significant improvement in their<br />

circulation time compared to that of free BSH. At 24 h following i.v. injection of PEG-liposomes,<br />

19% of the I.D. of boron was in the blood compared to 7% following formulation of BSH in<br />

conventional liposomes. The mean percent uptake <strong>by</strong> the liver and spleen was not significantly<br />

different for the two types of liposomes. However, the blood to RES ratios were higher for PEGliposomes<br />

at all time points indicating that a higher fraction of the injected dose (I.D.) of BSH was<br />

still in the blood. Ji et al. have reported that there were no significant differences in the in vitro<br />

uptake <strong>by</strong> 9L gliosarcoma cells of free BSH versus a liposomal formulation after 16 h incubation.<br />

However, cellular persistence was increased at 12 and 24 h for BSH-loaded liposomes. 111 BSH also<br />

has been incorporated into transferrin (TF) conjugated PEG liposomes (TF–PEG liposomes), 112<br />

which then were taken up <strong>by</strong> cells via TF receptor-mediated endocytosis. Intravenous administration<br />

of this formulation increased boron retention at the tumor site compared with PEG<br />

liposomes, bare liposomes or free BSH and suppressed tumor growth following BNCT. These<br />

results suggest that TF targeted liposomes might be useful as intracellular targeting vehicles.<br />

6.6.2.2 Liposomal Encapsulation of Other Boranes and Carboranes<br />

Polyhedral boranes 34 and carboranes 113,114 are another class of boron compounds that have been<br />

used for NCT. They contain multiple boron atoms per molecule, are resistant to metabolic<br />

degradation, and are lipophilic, there<strong>by</strong> permitting easier penetration of the tumor cell<br />

membrane. 113 In addition to BSH, carboranylpropylamine (CPA, Figure 6.7, 4) 115 has been incorporated<br />

into conventional and PEGylated liposomes <strong>by</strong> active loading, using a transmembrane pH<br />

gradient. 115,116 Although as many as 13,000 molecules of CPA were loaded into liposomes having a<br />

mean average diameter of 100 nm, there was no in vitro toxicity to both the glioblastoma cell line<br />

SK-MG-1 and normal human peripheral blood lymphocytes. Borane anions, such as B10H 2K<br />

10 ,<br />

B12H11SH 2K ,B20H17OH 4K , and B20H 3K<br />

19 , and the normal form and photoisomer of B20H 2K<br />

18 , also<br />

have been encapsulated into small unilamellar vesicles with mean diameters of 70 nm or less. 117–120<br />

These were composed of the pure synthetic phospholipids, distearoyl phosphatidylcholine, and<br />

cholesterol. 117 Although encapsulation efficiencies were only 2–3%, following i.v. injection,<br />

these liposomes were selectively delivered to the tumors in mice bearing the EMT6 mammary<br />

carcinoma and had attained boron concentrations of greater than 15 mg boron/g, with a T:Bl ratio of<br />

greater than 3. Two isomers of B 20H 2K<br />

18 attained boron concentrations of 13.6 and 13.9 mg/g with<br />

T:Bl ratios of 3.3 and 12, respectively, at 48 h following administration. High boron retention and<br />

prolongation of their circulation time was observed due to the interaction with intracellular components<br />

after it had been released from liposomes within tumor cells. 117<br />

To examine the effect of charge and substitution on the retention of boranes, two isomers<br />

[B 20H 17NH 3] 3K and [1-(2 0 -B 10H 9)-2-NH 3B 10H 8] 3K (5) were prepared from the polyhedral<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


88<br />

O O<br />

H<br />

4<br />

6<br />

8<br />

NH 3 + Cl –<br />

2–<br />

(CH 2 ) 15 CH 3<br />

–<br />

– +<br />

ClH3N(CH2 ) 3<br />

– +<br />

ClH3N(CH2 ) 4<br />

+Cl<br />

HN(CH2 ) 3<br />

–<br />

10 11<br />

Nanotechnology for Cancer Therapy<br />

borane anion [n-B 20H 18] 2K (6). 118,119 The sodium salts of these two isomers had been encapsulated<br />

within small unilamellar liposomes, composed of distearoyl phosphatidylcholine/cholesterol at a<br />

1:1 ratio. Both isomers of [B 20H 17NH 3] 3K had excellent tumor uptake and selectivity in EMT 6<br />

tumor bearing mice, even at very low I.D.s, and this resulted in peak tumor boron concentrations of<br />

30–40 mg B/g and a T:Bl ratio of approximately 5. Due to low boron retention of liposomal<br />

Na3[B20H19] andNa4[e 2 -B20H17OH] and rapid clearance of liposomal [2-NH3B10H9] K ,the<br />

enhanced retention of liposomal Na3[ae-B20H17NH3] was not due to the anionic charge or substitution<br />

in the borane cage. Rather, it could be attributed to their facile intracellular oxidation to an<br />

extremely reactive NH3-substituted [n-B20H18] 2K electrophilic anion, [B20H17NH3] K .Another<br />

anion [ae-B20H17NH3] 3K also was encapsulated into liposomes prepared with 5% PEG-2000distearoyl<br />

phosphatidylethanolamine as a constituent of the membrane. These liposomes had<br />

5<br />

7<br />

9<br />

NH 3<br />

SH<br />

H2 N<br />

3–<br />

4–<br />

NH 2<br />

+Cl –<br />

(CH 2 ) 3 NH 2<br />

3–<br />

BH<br />

B<br />

C<br />

CH<br />

NHCl –<br />

+<br />

FIGURE 6.7 Structures of hydrophilic and lipophilic boron-containing compounds that have been incorporated<br />

into liposomes. Carboranylpropylamine (CPA, 4), [1-(2 0 -B10H9)-2-NH3B10H8] 3K (5), [n-B20H18] 2K (6),<br />

[B 20H 17SH] 4K (7), Na 3[a2-B 20H 17NH 2CH 2CH 2NH 2](9) and boronated water soluble acridine (WSA, 11)<br />

were encapsulated into the aqueous core. K[nido-7-CH3(CH2)15-7,8-C2B9H11] (8) and cholesteryl 1,12dicarba-closo-dodecaboranel-carboxylate<br />

(10) were incorporated into the lipid bilayer of liposomes.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 89<br />

longer in vivo circulation times, which resulted in continued accumulation of boron in the tumor<br />

over the entire 48 h time period, and reached a maximum concentration of 47 mg B/g tumor.<br />

[B 20H 17SH] 4K (7), a thiol derivative of [B 20H 18] 4K , possesses a reactive thiol substituent and<br />

this can be oxidized into the more reactive [B 20H 17SH] 2K anion. Both of these were considered to<br />

be essential for high tumor boron retention 120 and they have been encapsulated into small, unilamellar<br />

liposomes. Biodistribution was determined after i.v. injection into BALB/c mice bearing<br />

EMT6 tumors. At low I.D.s, tumor boron concentrations increased throughout the duration of the<br />

experiment, resulting in a maximum concentration of 47 mg B/g tumor at 48 h, which corresponded<br />

to 22.2% I.D./g and a T:Bl ratio of 7.7. This was the most promising of the polyhedral borane anions<br />

that had been investigated for liposomal delivery. Although they were able to deliver adequate<br />

amounts of boron to tumor cells, their application to BNCT has been limited due to their low<br />

incorporation efficiency (approximately 3%).<br />

Lipophilic boron compounds incorporated into the lipid bilayer would be an alternative<br />

approach. Small unilamellar vesicles composed of 3:3:1 ratio of distearoylphosphatidylcholine,<br />

cholesterol and K[nido-7-CH3(CH2)15-7,8-C2B9H11] (8) in the lipid bilayer and Na3[a2-B20H17<br />

NH2CH2CH2NH2] (9) in the aqueous core were produced as a delivery agents for NCT mediated<br />

synovectomy. 121 Biodistribution studies were carried out in Louvain rats that had a collageninduced<br />

arthritis. The maximum synovial boron concentration was 29 mg/g tissue at 30 h and<br />

this had only decreased to 22 mg/g at 96 h following i.v. administration. The prolonged retention<br />

<strong>by</strong> synovium provided sufficient time for extensive clearance of boron from other tissues so that at<br />

96 h the synovium to blood (Syn:Bl) ratio was 3.0. To accelerate blood clearance, serum stability of<br />

the liposomes was lowered <strong>by</strong> increasing the proportion of K[nido-7-CH3(CH2)15-7,8-C2B9H11]<br />

embedded in the lipid bilayer. Liposomes were formulated with a 3:3:2 ratio of DSPC to Ch to<br />

K[nido-7-CH3(CH2)15-7,8-C2B9H11] in the lipid bilayer and Na3[a2-B20H17NH2CH2CH2NH2] was<br />

encapsulated in the aqueous core. The boron concentration in the synovium reached a maximum of<br />

26 mg/g at 48 h with a Syn:Bl ratio of 2, following which it slowly decreased to 14 mg/g at 96 h at<br />

which time the Syn:Bl ratio was 7.5. 121<br />

Another method to deliver hydrophilic boron containing compounds would be to incorporate<br />

them into cholesterol to target tumor cells expressing amplified low density lipoprotein (LDL)<br />

receptors. 122–124 Glioma cells, which absorb more cholesterol, have been reported to take up more<br />

LDL than the corresponding normal tissue cells. 125–127 The cellular uptake of liposomal cholesteryl<br />

1,12-dicarba-closo-dodecaboranel-carboxylate (10) <strong>by</strong> two fast growing human glioma cell lines,<br />

SF-763 and SF-767, was mediated via the LDL receptor and was much higher than that of human<br />

neurons. The cellular boron concentration was approximately 10–11 times greater than that<br />

required for BNCT. 128<br />

6.6.3 BORON DELIVERY BY TARGETED LIPOSOMES<br />

To improve the specificity of liposomally encapsulated drugs and to increase the amount of boron<br />

delivered, targeting moieties have been attached to the surface of liposomes. These could be any<br />

molecules that selectively recognized and bound to target antigens or receptors that were over<br />

expressed on neoplastic cells or tumor-associated neovasculature. These have included either intact<br />

mAb molecules or fragments, LMW, naturally occurring or synthetic ligands such as peptides or<br />

receptor-binding-ligands such as EGF. To date, liposomes linked to mAbs or their fragments, 129<br />

EGF, 130 folate, 131 and transferin, 112 have been the most extensively studied as targeting moieties<br />

(Figure 6.8).<br />

6.6.3.1 Immunoliposomes<br />

The murine anticarcinoembryonic antigen (CEA) mAb 2C-8 has been conjugated to large multilamellar<br />

liposomes containing 10 B compounds. 132,133 The maximum number of 10 B atoms attached<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


90<br />

A<br />

per molecule of mAb was approximately 1.2!10 4 . These immunoliposomes bound selectively to<br />

the human pancreatic carcinoma cell line, AsPC-1 that overexpressed CEA. Incubating the immunoliposomes<br />

with either MRKnu/nu-1 or AsPC-1 tumor cells, suppressed in vitro tumor cell growth<br />

following thermal neutron irradiation. 134 This was dependent upon the liposomal concentration of<br />

the 10 B-compound and on the number of molecules of mAb conjugated to the liposomes. Immunoliposomes<br />

containing either (Et4N)2B10H10 and linked to the mAb MGb 2, directed against<br />

human gastric cancer 135,136 or water-soluble boronated acridine (WSA, 11) linked to trastuzumab,<br />

directed against HER-2, have been prepared and evaluated in vitro. 129 There was specific binding<br />

and high uptake of these immunoliposomes, which delivered a sufficient amount of 10 B to produce a<br />

tumoricidal effect following thermal neutron irradiation.<br />

6.6.3.2 Folate Receptor-Targeted Liposomes<br />

C<br />

H +<br />

B<br />

A highly ionized boron compound, Na3B20H17NH3, was incorporated into liposomes <strong>by</strong> passive<br />

loading. 131,137,138 This showed high in vitro uptake <strong>by</strong> the FR expressing human cell line KB<br />

(American Type Culture Collection CCL 17), which originally was thought to be derived from a<br />

squamous cell carcinoma of the mouth, and subsequently was shown to be identical to HeLa cells,<br />

as determined <strong>by</strong> isoenyzyme markers, DNA fingerprinting, and karyotypic analysis. KB tumorbearing<br />

mice that received either FR-targeted or nontargeted control liposomes had equivalent<br />

tumor boron values (approximately 85 mg/g), which attained a maximum at 24 h, while the T:Bl<br />

ratio reached a maximum at 72 h. Additional studies were carried out with the lipophilic boron<br />

compound, K[nido-7-CH3(CH2)15-7,8-C2B9H11. This was incorporated into large unilamellar<br />

vesicles, approximately 200 nm in diameter, which were composed of egg PC/Chol/K[nido-7-<br />

CH3(CH2)15-7,8-C2B9H11] at a 2:2:1, mol/mol ratio, and an additional 0.5 mol% of folate–<br />

PEG–DSPE or PEG–DSPE for the FR-targeted or nontargeted liposomal formulations. 139 The<br />

boron uptake <strong>by</strong> FR-overexpressing KB cells, treated with these targeted liposomes, was approximately<br />

10 times greater compared with those treated with control liposomes. In addition, BSH and<br />

D<br />

H + H +<br />

Nanotechnology for Cancer Therapy<br />

Folate<br />

PEG<br />

Liposome<br />

Folate receptor<br />

Boron compound<br />

FIGURE 6.8 Proposed mechanism of intracellular boron delivery based on receptor-mediated tumor cell<br />

targeting. Liposomes loaded with a boron compound (A) are conjugated to a targeting ligand (e.g., folate,<br />

transferrin (TF), or anti-EGFR antibody). These bind to receptors on the cell surface (B), following which they<br />

are internalized <strong>by</strong> receptor mediated endocytosis (C) into the acidified endosomal/lysosomal compartment.<br />

The boron compound then is released (D) into the cytosol <strong>by</strong> liposomal degredation, endosome/lysosome<br />

disruption or liposome–endosome/lysome membrane fusion.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 91<br />

H 2 N<br />

H 2 N<br />

12<br />

N<br />

H 2 N N N H<br />

14 15<br />

N N<br />

16<br />

NH 2 · 3HCl<br />

NH 2 · 4HCl<br />

five weakly basic boronated polyamines were evaluated (Figure 6.9). Two of these were the<br />

spermidine derivatives N 5 -(4-carboranylbutyl)spermidine$3HCl (12) and N 5 -[4-(2-aminoethyl-ocarboranyl)butyl]<br />

spermidine$4HCl (13). Three were the spermine derivatives N 5 -(4-o-carboranylbutyl)<br />

spermine$4HCl (14), N 5 -[4-(2-aminoethyl-o-carboranyl) butyl] spermine$5HCl (15), and<br />

N 5 ,N 10 -bis(4-o-carboranylbutyl) spermine$4HCl (16). These were incorporated into liposomes<br />

<strong>by</strong> a pH-gradient-driven remote loading method with varying loading efficiencies that were influenced<br />

<strong>by</strong> the specific trapping agent and the structure of the boron compound. Greater loading<br />

efficiencies were obtained with lower molecular-weight boron derivatives, using ammonium sulfate<br />

as the trapping agent, compared to those obtained with sodium citrate.<br />

6.6.3.3 EGFR-Targeted Liposomes<br />

NH 2 · 4HCl<br />

Acridine is a WS, DNA-intercalator. Its boronated derivative WSA was incorporated into liposomes<br />

composed of EGF-conjugated lipids. Their surface contained approximately 5 mol% PEG<br />

and 10–15 molecules of EGF, and 10 4 –10 5 of WSA molecules were encapsulated. These liposomes<br />

had EGFR-specific cellular binding to cultured human glioma cells 130,140 and were internalized<br />

following specific binding to the receptor. Following internalization, WSA primarily was localized<br />

H 2 N<br />

H 2 N<br />

H 2 N · HCl<br />

H 2 N · HCl<br />

N<br />

13<br />

N<br />

BH<br />

C<br />

CH<br />

H<br />

N<br />

NH 2 · 3HCl<br />

NH 2 · 4HCl<br />

FIGURE 6.9 Structures of five weakly basic boronated polyamines encapsulated in FR-targeting liposomes.<br />

Two spermidine (12, 13) and three spermine derivatives (14–16) that contain hydrophilic amine groups and<br />

lipophilic carboranyl cages had DNA-binding properties.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


92<br />

in the cytoplasm, and had high cellular retention with 80% of the boron remaining cell-associated<br />

after 48 h. 141<br />

6.7 BORON DELIVERY BY DEXTRANS<br />

Dextrans are glucose polymers that consist mainly of a linear a-1,6-glucosidic linkage with some<br />

degree of branching via a 1,3-linkage. 142,143 Dextrans have been used extensively as drug and<br />

protein carriers to increase drug circulation time. 144,145 In addition, native or chemically-modified<br />

dextrans have been used for passive targeting to tumors, the RES or active receptor-specific cellular<br />

targeting. To link boron compounds to dextrans, 146 b-decachloro-o-carborane derivatives, in which<br />

one of the carbon atoms was substituted <strong>by</strong> –CH2CHOHCH2–O–CH2CHaCH2, were epoxidized<br />

and then subsequently bound to dextran with a resulting boron content of 4.3% (w/w). 147 The<br />

modified dextran then could be attached to tumor-specific antibodies. 147–150 BSH was covalently<br />

coupled to dextran derivatives <strong>by</strong> two methods. 151 In the first method, dextran was activated with<br />

1-cyano-4-(dimethylamino)pyridine (CDAP) and subsequently coupled with 2-aminoethyl pyridyl<br />

disulfide. Then, thiolated dextran was linked to BSH in a disulfide exchange reaction. A total of 10–<br />

20 boron cages were attached to each dextran chain. In the second method, dextran was derivatized<br />

to a multiallyl derivative (Figure 6.10, 17), which was reacted with BSH in a free-radical-initiated<br />

addition reaction. Using this method, 100–125 boron cages could be attached per dextran chain,<br />

suggesting that this derivative might be a promising template for the development of other HMW<br />

delivery agents. In the second method, designed to target EGFR overexpressing cells, EGF and<br />

BSH were covalently linked to a 70 kDa dextran (18). 152–154 Bioconjugates, having a small number<br />

of BSH molecules, attained maximum in vitro binding at 4 h with the human glioma cell line U-343<br />

O<br />

O<br />

O<br />

OH<br />

O<br />

OH<br />

O O<br />

OH<br />

O O<br />

OH O<br />

OH<br />

OH O<br />

OH O<br />

OH O<br />

OH O<br />

OH<br />

O O<br />

OH<br />

BSH or EGF-SH, 2 or 5h, 50 o C<br />

(NH 4 ) 2 S 2 O 8 /K 2 S 2 O 8<br />

O<br />

O<br />

O<br />

OH<br />

O<br />

OH<br />

O O<br />

OH<br />

O O<br />

S<br />

OH O S<br />

OH<br />

EGF<br />

OH O<br />

OH O<br />

17 18<br />

OH O<br />

OH O<br />

OH<br />

O O<br />

OH<br />

FIGURE 6.10 Preparation of EGF-targeted, boronated dextrans. The bioconjugate was prepared <strong>by</strong> a freeradical-initiated<br />

addition reaction between multiallyl dextran derivatives and BSH or thiolated EGF at 508C<br />

using K2S2O8 as an initiator.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

S<br />

BH<br />

B


Neutron Capture Therapy of Cancer 93<br />

MGaC12:6. In contrast, there was a slow increase of binding over 24 h for those having a large<br />

number of BSH molecules. Although most of the bioconjugates were internalized, in vitro retention<br />

was low, as was in vivo uptake following i.v. injection into nude mice bearing s.c. implants of<br />

Chinese hamster ovary (CHO) cells transfected with the human gene encoding EGFR (designated<br />

CHO–EGFR). However, following i.t. injection, boron uptake was higher with CHO–EGFR(C)<br />

tumors compared to wildtype EGFR(K) CHO tumors. 155<br />

6.8 OTHER MACROMOLECULES USED FOR DELIVERING<br />

BORON COMPOUNDS<br />

Polylysine is another polymer having multiple reactive amino groups that has been used as a<br />

platform for the delivery of boron compounds. 53,156 The protein-binding polyhedral boron derivatives,<br />

isocyanatoundecahydro-closo-dodecaborate (B12H11NCO 2K ), was linked to polylysine and<br />

subsequently to the anti-B16 melanoma mAb IB16-6 using two heterobifunctional linkers, SPDP<br />

and sulfo-MBS. The bioconjugate had an average of 2700 boron atoms per molecule and retained<br />

58% of the immunoreactivity of the native antibody, as determined <strong>by</strong> a semiquantitative immunofluorescent<br />

assay or <strong>by</strong> ELISA. Other bioconjugates prepared <strong>by</strong> this method had greater than 1000<br />

boron atoms per molecule of antibody and retained 40–90% of the immunoreactivity of the native<br />

antibody. 53 Using another approach, site-specific linkage of boronated polylysine to the carbohydrate<br />

moieties of anti-TSH antibody resulted in a bioconjugate that had approximately 6!10 3<br />

boron atoms with retention of its immunoreactivity. 156<br />

A streptavidin/biotin system also has been developed to specifically deliver boron to tumors.<br />

Biotin was linked to a mAb and streptavidin was attached to the boron containing moiety. The<br />

indirect linking of boron to the mAb minimized loss of its immunoreactivity. BSH was attached to<br />

poly-(D-glutamate D-lysine) (poly-GL) via a heterobifunctional agent. 157 This boronated poly-GL<br />

then was activated <strong>by</strong> a carbodiimide reagent and in turn reacted with streptavidin. Another<br />

approach employed a streptavidin mutant that had 20 cysteine residues per molecule. BSH was<br />

conjugated via sulfhydryl-specific bifunctional reagents to incorporate approximately 230 boron<br />

atoms/molecule. 158 A closomer species with an icosahedral dodecaborate core and twelve pendant<br />

anionic nido-7,8-carborane groups was developed as a new class of unimolecular nanovehicles for<br />

evaluation as a delivery agent for BNCT. 159<br />

6.9 DELIVERY OF BORON-CONTAINING MACROMOLECULES<br />

TO BRAIN TUMORS<br />

6.9.1 GENERAL CONSIDERATIONS<br />

Drug delivery to brain tumors is dependent upon (1) the plasma concentration profile of the drug,<br />

which depends upon the amount and route of administration; (2) the ability of the agent to traverse<br />

the BBB; (3) blood flow within the tumor; and (4) the lipophilicity of the drug. In general, a high<br />

steady-state blood concentration will maximize brain uptake, while rapid clearance will reduce it,<br />

except in the case of intra-arterial (i.a.) drug administration. Although the i.v. route currently is<br />

being used clinically to administer both BSH and BPA, this may not be ideal for boron containing<br />

macromolecules and other strategies must be employed to improve their delivery.<br />

6.9.2 DRUG-TRANSPORT VECTORS<br />

One approach to improve brain tumor uptake of boron compounds has been to conjugate them to a<br />

drug-transport vector <strong>by</strong> means of receptor-specific transport systems. 160,161 Proteins such as<br />

insulin, insulin-like growth factor (IGF), TF, 162 and leptin can traverse the BBB. BSH encapsulated<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


94<br />

in TF–PEG liposomes had a prolonged residence time in the circulation and low RES uptake in<br />

tumor-bearing mice, resulting in enhanced extravasation of the liposomes into the tumor and<br />

concomitant internalization <strong>by</strong> receptor-mediated endocytosis. 163,164 Mice that received BSH<br />

containing TF-liposomes following <strong>by</strong> BNCT had a significant prolongation in survival time<br />

compared to those that received PEG-liposomes, bare liposomes and free BSH, there<strong>by</strong> establishing<br />

proof-of-principle for transcytosis of a boron containing nanovehicle. 112<br />

6.9.3 DIRECT INTRACEREBRAL DELIVERY<br />

Studies carried out <strong>by</strong> us have clearly demonstrated that the systemic route of administration is not<br />

suitable for delivery of boronated EGF or mAbs to glioma-bearing rats. 75,165 Intravenous injection<br />

of technetium-99m labeled EGF to rats bearing intracerebral implants of the C6 rat glioma, which<br />

had been genetically engineered to express the human EGFR gene, resulted in 0.14% I.D. localizing<br />

in the tumor. Intracarotid (i.c.) injection with or without BBB disruption increased the tumor uptake<br />

from 0.34 to 0.45% I.D./g, but based even on the most optimistic assumptions the amount of boron<br />

that could be delivered to the tumor <strong>by</strong> i.v. injection, this would have been inadequate for BNCT. 165<br />

Direct i.t. injection of boronated EGF (BSD–EGF), on the other hand, resulted in tumor boron<br />

concentrations of 22 mg/g compared to 0.01 mg/g following i.v. injection and almost identical boron<br />

uptake values were obtained using the F98 EGFR glioma model. 77 This was produced <strong>by</strong> transfecting<br />

F98 glioma cells with the gene encoding human EGFR. Based on our biodistribution results,<br />

therapy studies were initiated with the F98EGFR glioma in syngeneic Fischer rats. F98EGFR<br />

glioma bearing rats that received BSD–EGF i.t. had a MST of 45Gd compared to 33G2 din<br />

animals that had EGFR(K) wildtype F98 gliomas. Because it is unlikely that any single boron<br />

delivery agent will be able to target all tumor cells, the combination of i.t. administration of<br />

BSD–EGF with i.v. injection of BPA was evaluated. This resulted in further increase in MST to<br />

57G8 d compared to 39G2 d for i.v. BPA alone. 73 These data provide proof-of-principle for the<br />

idea of using a combination of LMW and HMW boron delivery agents.<br />

6.9.4 CONVECTION-ENHANCED DELIVERY<br />

Nanotechnology for Cancer Therapy<br />

Convection-Enhanced Delivery (CED), <strong>by</strong> which therapeutic agents are directly infused into the<br />

brain, is an innovative method to increase their uptake and distribution. 166–168 Under normal<br />

physiological conditions, interstitial fluids move through the brain <strong>by</strong> both convection and diffusion.<br />

Diffusion of a drug in tissue depends upon its molecular weight, ionic charge and its<br />

concentration gradient within normal tissue and the tumor. The higher the molecular weight of<br />

the drug, the more positively charged the ionic species and the lower its concentration, the slower<br />

its diffusion. For example, diffusion of antibody into a tumor requires 3 d to diffuse 1 mm from the<br />

point of origin. Unlike diffusion, however, convection or “bulk” flow results from a pressure<br />

gradient that is independent of themolecularweightofthesubstance. CED potentially can<br />

improve the targeting of both LMW and HMW molecules, as well as liposomes, to the central<br />

nervous system <strong>by</strong> applying a pressure gradient to establish bulk flow during interstitial infusion.<br />

The volume of distribution (V d) is a linear function of the volume of the infusate (V i). CED has been<br />

used to efficiently deliver drugs and HMW agents such as mAbs and toxin fusion proteins to brain<br />

tumors. 168–170 CED can provide more homogenous dispersion of the agent and at higher concentrations<br />

than otherwise would be attainable <strong>by</strong> i.v. injection. 165 For example, in our own studies,<br />

CED of 125 I-labeled EGF to F98EGFR glioma-bearing rats resulted in 47% I.D./g of the bioconjugate<br />

localizing in the tumor compared to 10% I.D./g in normal brain at 24 h following administration.<br />

The corresponding boron values were 22 and 2.9–4.9 mg/g, respectively. 76 Based on these results,<br />

therapy studies were initiated. F98EGFR glioma-bearing rats that received BD–EGF <strong>by</strong> CED had a<br />

MST of 53G13 d compared to 40G5 d for animals that received BPA i.v. 73 Similar studies have<br />

been carried out using either boronated cetuximab (IMC-C225) or the mAb L8A4, 171,172 which is<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 95<br />

specifically directed against the tumor-specific mutant isoform, EGFRvIII, and comparable results<br />

were obtained. 173 Direct intracerebral administration of these and other HMW agents <strong>by</strong> CED has<br />

opened up the possibility that they actually could be used clinically, since CED is being used to<br />

administer radiolabeled antibodies, toxin fusion proteins, and gene vectors to patients with GBM. It<br />

is only a matter of time before this approach also will be used to deliver both LMW and HMW<br />

boron-containing agents for NCT.<br />

6.10 CLINICAL CONSIDERATIONS AND CONCLUSIONS<br />

In this review we have focused on HMW boron delivery agents and nanovehicles that potentially<br />

could be used clinically for targeting intra- and extracranial tumors. Studies carried out in glioma<br />

bearing rats have demonstrated that boronated EGF and the mAb, cetuximab, both of which bind to<br />

EGFR, selectively targeted receptor (C) tumors following direct i.c. delivery. Furthermore,<br />

following BNCT, a significant increase in MST was observed, and this was further enhanced if<br />

BPA was administered in combination with the HMW agents. These studies provide proof-of-principle<br />

first for the potential utility of HMW agents, and second, the therapeutic gain associated with<br />

the combination of HMW and LMW boron delivery agents.<br />

There is a question as to whether or not any of these agents will ever be used clinically. Critical<br />

issues must be addressed if BNCT is to ever become a useful modality for the treatment of cancer.<br />

Large clinical trials, preferably randomized, must be carried out to convincingly demonstrate the<br />

efficacy of BPA and BSH, the two drugs that currently are being used. Once efficacy has been<br />

established, studies with HMW EGFR targeting agents could move forward.<br />

Both direct i.t. injection 174,175 and CED 170,176–179 have been used clinically to deliver mAbs<br />

and toxin fusion proteins to patients who have had surgical resection of their brain tumors. These<br />

studies provide a strong clinical rationale for the direct intracerabral delivery of HMW agents.<br />

Initially, the primary focus should be on determining the safety of administering them to patients<br />

prior to surgical resection of their brain tumors. After this has been established, then biodistribution<br />

studies could be carried out in patients scheduled to undergo surgical resection of their brain<br />

tumors. The patients’ tumor and normal tissues would then be analyzed for their boron content,<br />

and if there was evidence of preferential tumor localization with boron concentrations in the range<br />

10–20 mg/g and normal brain concentrations of less than 5 mg/g, then therapy studies could<br />

be undertaken.<br />

Because there is considerable variability in EGFR expression in gliomas, it is highly unlikely<br />

that any single agent will be able to deliver the requisite amount of boron to all tumor cells, and that<br />

HMW agents would need to be used in combination with BPA/BSH. This general plan would also<br />

be applicable to the other HMW delivery agents and nanovehicles that have been discussed in this<br />

review. The joining together of chemistry and nanotechnology 180,181 represents a major step<br />

forward for the development of effective boron delivery agents for NCT. Nanovehicles offer the<br />

possibility of tumor targeting with enhanced boron payloads. Potentially, this could solve the<br />

central problem of how to selectively deliver large number of boron atoms to individual<br />

cancer cells.<br />

As can be seen from the preceding discussion, the development of HMW boron delivery agents<br />

must proceed in step with strategies to optimize their delivery and an appreciation as to how they<br />

would be used clinically. Intracerebral delivery has been used in clinically advanced settings, but<br />

nuclear reactors, which currently are the only source of neutrons for BNCT, would not be conducive<br />

to this. Therefore, the development of accelerator-neutron sources, 182 which could be easily<br />

sited in hospitals, is especially important. This also would facilitate the initiation of large scale<br />

clinical trials at selected centers that treat large numbers of patients with brain tumors and would<br />

permit evaluation of new boron delivery agents.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


96<br />

In conclusion, there is a plethora of HMW boron delivery agents that have been designed and<br />

synthesized. The challenge is to move from experimental animal studies to clinical biodistribution<br />

studies, a step which has yet to be taken.<br />

6.11 SUMMARY<br />

BNCT is based on the nuclear capture and fission reactions that occur when nonradioactive boron-<br />

10 is irradiated with low energy thermal neutrons to yield LET alpha particles ( 4 He) and recoiling<br />

lithium-7 ( 7 Li) nuclei. For BNCT to be successful, a sufficient number of 10 B atoms (approximately<br />

10 9 atoms/cell) must be selectively delivered to the tumor and enough thermal neutrons must be<br />

absorbed <strong>by</strong> them to sustain a lethal 10 B(n,a) 7 Li capture reaction. BNCT primarily has been used to<br />

treat patients with brain tumors, and more recently those with head-and-neck cancer. Two LMW<br />

boron delivery agents currently are being used clinically, BSH and BPA. However, a variety of<br />

HMW agents consisting of macromolecules and nanovehicles have been developed. This review<br />

focuses on the latter, which includes mAbs, dendrimers, liposomes, dextrans, polylysine, avidin,<br />

FA, and both epidermal and vascular endothelial growth factors (EGF and VEGF). Procedures for<br />

introducing boron atoms into these HMW agents and their chemical properties are discussed.<br />

In vivo studies on their biodistribution are described, and the efficacy of a subset of them, those<br />

which have been used for BNCT of tumors in experimental animals, will be discussed.<br />

Because brain tumors currently are the primary candidates for treatment <strong>by</strong> BNCT, delivery of<br />

these HMW agents across the BBB presents a special challenge. Various routes of administration are<br />

discussed, including receptor-facilitated transcytosis following i.v. administration, direct i.t injection<br />

and convection-enhanced delivery in which a pump is used to apply a pressure gradient to establish<br />

bulk flow of the HMW agent during interstitial infusion. Finally, we have concluded with a discussion<br />

relating to issues that must be addressed if these HMW agents are to be used clinically.<br />

ACKNOWLEDGMENTS<br />

We thank Dr. Achintya Bandyopadhyaya for suggestions on figures and Mrs. Beth Kahl for<br />

secretarial assistance. We thank Bentham Science Publishers, Ltd. for granting copyright release<br />

to republish the text and figures in this chapter.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Berger, M. S., Malignant astrocytomas: Surgical aspects, Semin. Oncol., 21, 172, 1994.<br />

2. Gutin, P. H. and Posner, J. B., Neuro-oncology: Diagnosis and management of cerebral gliomas—<br />

past, present, and future, Neurosurgery, 47, 1, 2000.<br />

3. Parney, I. F. and Chang, S. M., Current chemotherapy for glioblastoma, Cancer J., 9, 149, 2003.<br />

4. Paul, D. B. and Kruse, C. A., Immunologic approaches to therapy for brain tumors, Curr. Neurol.<br />

Neurosci. Rep., 1, 238, 2001.<br />

5. Rainov, N. G. and Ren, H., Gene therapy for human malignant brain tumors, Cancer J., 9, 180, 2003.<br />

6. Curran, W.J., Jr., et al., Recursive partitioning analysis of prognostic factors in three Radiation<br />

Therapy Oncology <strong>Group</strong> malignant glioma trials, J. Natl Cancer Inst., 85, 704, 1993.<br />

7. Lacroix, M. et al., A multivariate analysis of 416 patients with glioblastoma multiforme: Prognosis,<br />

extent of resection, and survival, J. Neurosurg., 95, 190, 2001.<br />

8. Hentschel, S. J. and Lang, F. F., Current surgical management of glioblastoma, Cancer J., 9, 113,<br />

2003.<br />

9. Laws, E.R., Jr. and Shaffrey, M. E., The inherent invasiveness of cerebral gliomas: Implications for<br />

clinical management, Int. J. Dev. NeuroSci., 17, 413, 1999.<br />

10. Ware, M. L., Berger, M. S., and Binder, D. K., Molecular biology of glioma tumorigenesis, Histol.<br />

Histopathol., 18, 207, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 97<br />

11. Parney, I. F., Hao, C., and Petruk, K. C., Glioma immunology and immunotherapy, Neurosurgery,<br />

46, 778, 2000.<br />

12. Kaczarek, E. et al., Dissecting glioma invasion: Interrelation of adhesion, migration and intercellular<br />

contacts determine the invasive phenotype, Int. J. Dev. NeuroSci., 17, 625, 1999.<br />

13. Nutt, C. L., Matthews, R. T., and Hockfield, S., Glial tumor invasion: A role for the upregulation and<br />

cleavage of BEHAB/brevican, Neuroscientist, 7, 113, 2001.<br />

14. Halperin, E. C., Burger, P. C., and Bullard, D. E., The fallacy of the localized supratentorial<br />

malignant glioma, Int. J. Radiat. Oncol. Biol. Phys., 15, 505, 1988.<br />

15. Barth, R. F., A critical assessment of boron neutron capture therapy: An overview, J. Neurooncol.,<br />

62, 1, 2003.<br />

16. Nakagawa, Y. et al., Clinical review of the Japanese experience with boron neutron capture therapy<br />

and a proposed strategy using epithermal neutron beams, J. Neurooncol., 62, 87, 2003.<br />

17. Wadabayashi, N. et al., Selective boron accumulation in human ocular melanoma vs. surrounding<br />

eye components after 10 B1-p-boronophenylalanine administration. Prerequisite for clinical trial of<br />

neutron-capture therapy, Melanoma Res., 4, 185, 1994.<br />

18. Busse, P. M. et al., A critical examination of the results from the Harvard-MIT NCT program phase I<br />

clinical trial of neutron capture therapy for intracranial disease, J. Neurooncol., 62, 111, 2003.<br />

19. Kato, I. et al., Effectiveness of BNCT for recurrent head and neck malignancies, Appl. Radiat. Isot.,<br />

61, 1069, 2004.<br />

20. Rao, M. et al., BNCT of 3 cases of spontaneous head and neck cancer in feline patients, Appl. Radiat.<br />

Isot., 61, 947, 2004.<br />

21. Koivunoro, H. et al., BNCT dose distribution in liver with epithermal D–D and D–T fusion-based<br />

neutron beams, Appl. Radiat. Isot., 61, 853, 2004.<br />

22. Pinelli, T. et al., TAOrMINA: from the first idea to the application to the human liver, In Research<br />

and Development in Neutron Capture Therapy, Sauerwein, M. W., Moss, R., Wittig, A. et al., Eds.,<br />

Modduzzi Editore, International Proceedings Division, Bologna, pp. 1065–1072, 2002.<br />

23. Coderre, J. A. et al., Boron neutron capture therapy: Cellular targeting of high linear energy transfer<br />

radiation, Technol. Cancer Res. Treat., 2, 355, 2003.<br />

24. Barth, R. F. et al., Boron neutron capture therapy of cancer: Current status and future prospects, Clin.<br />

Cancer Res., 11, 3987, 2005.<br />

25. Zamenhof, R. G. et al., Eleventh World Congress on Neutron Capture Therapy, Appl. Radiat. Isot.,<br />

61, 731, 2004.<br />

26. Farr, L. E. et al., Neutron capture therapy with boron in the treatment of glioblastoma multiforme,<br />

Am. J. Roentgenol. Radium Ther. Nucl. Med., 71, 279, 1954.<br />

27. Goodwin, J. T. et al., Pathological study of eight patients with glioblastoma multiforme treated <strong>by</strong><br />

neutron-capture therapy using boron 10, Cancer, 8, 601, 1955.<br />

28. Snyder, H. R., Reedy, A. J., and Lennarz, W. J., Synthesis of aromatic boronic acids. Aldehydo<br />

boronic acids and a boronic acid analog of tyrosine, J. Am. Chem. Soc., 80, 835, 1958.<br />

29. Soloway, A. H., Hatanaka, H., and Davis, M. A., Penetration of brain and brain tumor. VII. Tumorbinding<br />

sulfhydryl boron compounds, J. Med. Chem., 10, 714, 1967.<br />

30. Hatanaka, H. and Nakagawa, Y., Clinical results of long-surviving brain tumor patients who underwent<br />

boron neutron capture therapy, Int. J. Radiat. Oncol. Biol. Phys., 28, 1061, 1994.<br />

31. Mishima, Y., Selective thermal neutron capture therapy of cancer cells using their specific metabolic<br />

activities—melanoma as prototype, In Cancer Neutron Capture Therapy, Mishima, Y., Ed., Plenum<br />

Press, New York, pp. 1–26, 1996.<br />

32. Ang, K. K. et al., Impact of epidermal growth factor receptor expression on survival and pattern of<br />

relapse in patients with advanced head and neck carcinoma, Cancer Res., 62, 7350, 2002.<br />

33. Hawthorne, M. F. and Lee, M. W., A critical assessment of boron target compounds for boron<br />

neutron capture therapy, J. Neurooncol., 62, 33, 2003.<br />

34. Soloway, A. H. et al., The chemistry of neutron capture therapy, Chem. Rev., 98, 1515, 1998.<br />

35. Gillies, E. R. and Frechet, J. M. J., Dendrimers and dendritic polymers in drug delivery, Drug<br />

Discov. Today, 10, 35, 2005.<br />

36. Esfand, R. and Tomalia, D. A., Poly(amidoamine) (PAMAM) dendrimers: From biomimicry to drug<br />

delivery and biomedical applications, Drug Discov. Today, 6, 427, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


98<br />

Nanotechnology for Cancer Therapy<br />

37. Tomalia, D.A., Birth of a new macromolecular architecture: Dendrimers as quantized building<br />

blocks for nanoscale synthetic organic chemistry, Aldrichimica ACTA, 37, 39, 2004<br />

38. Verheyde, B., Maes, W., and Dehaen, W., The use of 1,3,5-triazines in dendrimer synthesis, Mater.<br />

Sci. Eng., C, 18, 243, 2001.<br />

39. McCarthy, T. D. et al., Dendrimers as drugs: Discovery and preclinical and clinical development of<br />

dendrimer-based microbicides for HIV and STI prevention, Mol. Pharmacol., 2, 312, 2005.<br />

40. Venditto, V. J., Regino, C. A., and Brechbiel, M. W., PAMAM dendrimer based macromolecules as<br />

improved contrast agents, Mol. Pharmacol., 2, 302, 2005.<br />

41. Ambade, A. V., Savariar, E. N., and Thayumanavan, S., Dendrimeric micelles for controlled drug<br />

release and targeted delivery, Mol. Pharmacol., 2, 264, 2005.<br />

42. Majoros, I. J. et al., Poly(amidoamine) dendrimer-based multifunctional engineered nanodevice for<br />

cancer therapy, J. Med. Chem., 48, 5892, 2005.<br />

43. Boas, U. and Heegaard, P. M., Dendrimers in drug research, Chem. Soc. Rev., 33, 43, 2004.<br />

44. Klajnert, B. and Bryszewska, M., Dendrimers: Properties and applications, Acta Biochim. Pol., 48,<br />

199, 2001.<br />

45. Sharkey, R. M. and Goldenberg, D. M., Perspectives on cancer therapy with radiolabeled monoclonal<br />

antibodies, J. Nucl. Med., 46(Suppl 1), 115S, 2005.<br />

46. Jaracz, S. et al., Recent advances in tumor-targeting anticancer drug conjugates, Bioorg. Med.<br />

Chem., 13, 5043, 2005.<br />

47. Garnett, M. C., Targeted drug conjugates: Principles and progress, Adv. Drug Deliv. Rev., 53, 171,<br />

2001.<br />

48. Chari, R. V., Targeted delivery of chemotherapeutics: Tumor-activated prodrug therapy, Adv. Drug<br />

Deliv. Rev., 31, 89, 1998.<br />

49. Hamblett, K. J. et al., Effects of drug loading on the antitumor activity of a monoclonal antibody drug<br />

conjugate, Clin. Cancer Res., 10, 7063, 2004.<br />

50. Trail, P. A., King, H. D., and Dubowchik, G. M., Monoclonal antibody drug immunoconjugates for<br />

targeted treatment of cancer, Cancer Immunol. Immunother., 52, 328, 2003.<br />

51. Fracasso, G. et al., Immunotoxins and other conjugates: Preparation and general characteristics,<br />

Mini-Rev. Med. Chem., 4, 545, 2004.<br />

52. Liu, L. et al., Bispecific antibodies as targeting agents for boron neutron capture therapy of brain<br />

tumors, J. Hematother., 4, 477, 1995.<br />

53. Alam, F. et al., Boron neutron capture therapy: Linkage of a boronated macromolecule to monoclonal<br />

antibodies directed against tumor-associated antigens, J. Med. Chem., 32, 2326, 1989.<br />

54. Barth, R. F. et al., Neutron capture using boronated monoclonal antibody directed against tumorassociated<br />

antigens, Cancer Detect. Prev., 5, 315, 1982.<br />

55. Alam, F., Barth, R. F., and Soloway, A. H., Boron containing immunoconjugates for neutron capture<br />

therapy of cancer and for immunocyto chemistry, Antibody Immunoconjugates Radiopharm., 2, 145,<br />

1989.<br />

56. Tolpin, E. I. et al., Boron neutron capture therapy of cerebral gliomas. II. Utilization of the blood–<br />

brain barrier and tumor-specific antigens for the selective concentration of boron in gliomas,<br />

Oncology, 32, 223, 1975.<br />

57. Sneath, R. L. et al., Protein-binding polyhedral boranes, J. Med. Chem., 19, 1290, 1976.<br />

58. Varadarajan, A. et al., Conjugation of phenyl isothiocyanate derivatives of carborane to antitumor<br />

antibody and in vivo localization of conjugates in nude mice, Bioconjug. Chem., 2, 102, 1991.<br />

59. Takahashi, T. et al., Preliminary study for application of anti-alpha-fetoprotein monoclonal antibody<br />

to boron-neutron capture therapy, Jpn. J. Exp. Med., 57, 83, 1987.<br />

60. Compostella, F. et al., Synthesis of glycosyl carboranes with different linkers between the sugar and<br />

the boron cage moieties, In Research and Development in Neutron Capture Therapy. Proceedings of<br />

the 10th International Congress on Neutron Capture Therapy, Saverwein, W., Moss, R., and Wittig,<br />

A., Eds., Monduzzi Editore, Bologna, p.8, 2002.<br />

61. Giovenzana, G. B. et al., Synthesis of carboranyl derivatives of alkynyl glycosides as potential<br />

BNCT agents, Tetrahedron, 55, 14123, 1999.<br />

62. Alam, F. et al., Dicesium N-succinimidyl 3-(undecahydro-closo-dodecaboranyldithio)propionate, a<br />

novel heterobifunctional boronating agent, J. Med. Chem., 28, 522, 1985.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 99<br />

63. Barth, R. F. et al., Boronated starburst dendrimer-monoclonal antibody immunoconjugates: evaluation<br />

as a potential delivery system for neutron capture therapy, Bioconjug. Chem., 5, 58, 1994.<br />

64. Liu, L. et al., Critical evaluation of bispecific antibodies as targeting agents for boron neutron capture<br />

therapy of brain tumors, Anticancer Res., 16, 2581, 1996.<br />

65. Barth, R. F. et al., In vivo distribution of boronated monoclonal antibodies and starburst dendrimers,<br />

In Adv Neutron Capture Ther, Soloway, A. H., Barth, R. F., and Carpenter, D.E., Eds., Plenum Press,<br />

New York, p. 35, 1993.<br />

66. Wu, G. et al., Site-specific conjugation of boron-containing dendrimers to anti-EGF receptor monoclonal<br />

antibody cetuximab (IMC-C225) and its evaluation as a potential delivery agent for neutron<br />

capture therapy, Bioconjug. Chem., 15, 185, 2004.<br />

67. Barth, R. F. et al., Neutron capture therapy of epidermal growth factor (C) gliomas using boronated<br />

cetuximab (IMC-C225) as a delivery agent, Appl. Radiat. Isot., 61, 899, 2004.<br />

68. Arteaga, C. L., Overview of epidermal growth factor receptor biology and its role as a therapeutic<br />

target in human neoplasia, Semin. Oncol., 29, 3, 2002.<br />

69. Mendelson, J. and Beselga, J., Epidermal growth factor receptor targeting in cancer, Seminars<br />

Oncol., 33, 369, <strong>2006</strong>.<br />

70. Pal, S. K. and Pegram, M., Epidermal growth factor receptor and signal transduction: potential<br />

targets for anti-cancer therapy, Anticancer Drugs, 16, 483, 2005.<br />

71. Normanno, N. et al., The ErbB receptors and their ligands in cancer: an overview, Curr. Drug<br />

Targets, 6, 243, 2005.<br />

72. Jorissen, R. N. et al., Epidermal growth factor receptor: mechanisms of activation and signaling, Exp.<br />

Cell Res., 284, 31, 2003.<br />

73. Yang, W. et al., Boronated epidermal growth factor as a delivery agent for neutron capture therapy of<br />

EGF receptor positive gliomas, Appl. Radiat. Isot., 61, 981, 2004.<br />

74. Capala, J. et al., Boronated epidermal growth factor as a potential targeting agent for boron neutron<br />

capture therapy of brain tumors, Bioconjug. Chem., 7, 7, 1996.<br />

75. Yang, W. et al., Intratumoral delivery of boronated epidermal growth factor for neutron capture<br />

therapy of brain tumors, Cancer Res., 57, 4333, 1997.<br />

76. Yang, W. et al., Convection-enhanced delivery of boronated epidermal growth factor for molecular<br />

targeting of EGF receptor-positive gliomas, Cancer Res., 62, 6552, 2002.<br />

77. Barth, R. F. et al., Molecular targeting of the epidermal growth factor receptor for neutron capture<br />

therapy of gliomas, Cancer Res., 62, 3159, 2002.<br />

78. Reddy, J. A., Allagadda, V. M., and Leamon, C. P., Targeting therapeutic and imaging agents to<br />

folate receptor positive tumors, Curr. Pharm. Biotechnol., 6, 131, 2005.<br />

79. Leamon, C. P. and Reddy, J. A., Folate-targeted chemotherapy, Adv. Drug Deliv. Rev., 56, 1127,<br />

2004.<br />

80. Sudimack, J. and Lee, R. J., Targeted drug delivery via the folate receptor, Adv. Drug Deliv. Rev., 41,<br />

147, 2000.<br />

81. Paulos, C. M. et al., Ligand binding and kinetics of folate receptor recycling in vivo: impact on<br />

receptor-mediated drug delivery, Mol. Pharmacol., 66, 1406, 2004.<br />

82. Reddy, J. A. and Low, P. S., Folate-mediated targeting of therapeutic and imaging agents to cancers,<br />

Crit. Rev. Ther. Drug Carrier Syst., 15, 587, 1998.<br />

83. Stephenson Stacy, M. et al., Folate receptor-targeted liposomes as possible delivery vehicles for<br />

boron neutron capture therapy, Anticancer Res., 23, 3341, 2003.<br />

84. Ward, C. M., Folate-targeted non-viral DNA vectors for cancer gene therapy, Curr. Opin. Mol. Ther.,<br />

2, 182, 2000.<br />

85. Gottschalk, S. et al., Folate receptor mediated DNA delivery into tumor cells: Protosomal disruption<br />

results in enhanced gene expression, Gene Ther., 1, 185, 1994.<br />

86. Hilgenbrink, A. R. and Low, P. S., Folate receptor-mediated drug targeting: from therapeutics to<br />

diagnostics, J. Pharm. Sci., 94, 2135, 2005.<br />

87. Shukla, S. et al., Synthesis and biological evaluation of folate receptor-targeted boronated PAMAM<br />

dendrimers as potential agents for neutron capture therapy, Bioconjug. Chem., 14, 158, 2003.<br />

88. Benouchan, M. and Colombo, B. M., Anti-angiogenic strategies for cancer therapy (Review), Int.<br />

J. Oncol., 27, 563, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


100<br />

Nanotechnology for Cancer Therapy<br />

89. Tortora, G., Melisi, D., and Ciardiello, F., Angiogenesis: a target for cancer therapy, Curr. Pharm.<br />

Des., 10, 11, 2004.<br />

90. Brekken, R. A., Li, C., and Kumar, S., Strategies for vascular targeting in tumors, Int. J. Cancer, 100,<br />

123, 2002.<br />

91. Gaya, A. M. and Rustin, G. J., Vascular disrupting agents: a new class of drug in cancer therapy,<br />

Clin. Oncol. (R. Coll. Radiol.), 17, 277, 2005.<br />

92. Bergsland, E. K., Vascular endothelial growth factor as a therapeutic target in cancer, Am. J. Health<br />

Syst. Pharm., 61, S4, 2004.<br />

93. Hicklin, D. J. and Ellis, L. M., Role of the vascular endothelial growth factor pathway in tumor<br />

growth and angiogenesis, J. Clin. Oncol., 23, 1011, 2005.<br />

94. Backer, M. V. et al., Vascular endothelial growth factor selectively targets boronated dendrimers to<br />

tumor vasculature, Mol. Cancer Ther., 4, 1423, 2005.<br />

95. Barth, R. F., Yang, W., and Coderre, J. A., Rat brain tumor models to assess the efficacy of boron<br />

neutron capture therapy: a critical evaluation, J. Neurooncol., 62, 61, 2003.<br />

96. Parrott, M. C. et al., Synthesis and properties of carborane-functionalized aliphatic polyester dendrimers,<br />

J. Am. Chem. Soc., 127, 12081, 2005.<br />

97. Park, J. W., Benz, C. C., and Martin, F. J., Future directions of liposome- and immunoliposomebased<br />

cancer therapeutics, Semin. Oncol., 31, 196, 2004.<br />

98. Sapra, P. and Allen, T. M., Ligand-targeted liposomal anticancer drugs, Prog. Lipid Res., 42, 439,<br />

2003.<br />

99. Carlsson, J. et al., Ligand liposomes and boron neutron capture therapy, J. Neurooncol., 62, 47, 2003.<br />

100. Hawthorne, M. F. and Shelly, K., Liposomes as drug delivery vehicles for boron agents,<br />

J. Neurooncol., 33, 53, 1997.<br />

101. Mishima, Y. et al., Treatment of malignant melanoma <strong>by</strong> single thermal neutron capture therapy with<br />

melanoma-seeking 10 B-compound, Lancet, 2, 388, 1989.<br />

102. Coderre, J. A. et al., Selective delivery of boron <strong>by</strong> the melanin precursor analogue p-boronophenylalanine<br />

to tumors other than melanoma, Cancer Res., 50, 138, 1990.<br />

103. Ono, K. et al., The combined effect of boronophenylalanine and borocaptate in boron neutron capture<br />

therapy for SCCVII tumors in mice, Int. J. Radiat. Oncol. Biol. Phys., 43, 431, 1999.<br />

104. Obayashi, S. et al., Delivery of (10)boron to oral squamous cell carcinoma using boronophenylalanine<br />

and borocaptate sodium for boron neutron capture therapy, Oral Oncol., 40, 474, 2004.<br />

105. Yoshino, K. et al., Improvement of solubility of p-boronophenylalanine <strong>by</strong> complex formation with<br />

monosaccharides, Strahlenther. Onkol., 165, 127, 1989.<br />

106. Ryynanen, P. M. et al., Models for estimation of the 10 B concentration after BPA-fructose complex<br />

infusion in patients during epithermal neutron irradiation in BNCT, Int. J. Radiat. Oncol. Biol. Phys.,<br />

48, 1145, 2000.<br />

107. Pavanetto, F. et al., Boron-loaded liposomes in the treatment of hepatic metastases: preliminary<br />

investigation <strong>by</strong> autoradiography analysis, Drug Deliv., 7, 97, 2000.<br />

108. Martini, S. et al., Boronphenylalanine insertion in cationic liposomes for boron neutron capture<br />

therapy, Biophys. Chem., 111, 27, 2004.<br />

109. Smyth Templeton, N. et al., Cationic liposomes as in vivo delivery vehicles, Curr. Med. Chem., 10,<br />

1279, 2003.<br />

110. Mehta, S. C., Lai, J. C., and Lu, D. R., Liposomal formulations containing sodium mercaptoundecahydrododecaborate<br />

(BSH) for boron neutron capture therapy, J. Microencapsul., 13, 269, 1996.<br />

111. Ji, B. et al., Cell culture and animal studies for intracerebral delivery of borocaptate in liposomal<br />

formulation, Drug Deliv., 8, 13, 2001.<br />

112. Maruyama, K. et al., Intracellular targeting of sodium mercaptoundecahydrododecaborate (BSH) to<br />

solid tumors <strong>by</strong> transferrin-PEG liposomes, for boron neutron-capture therapy (BNCT), J. Control.<br />

Release, 98, 195, 2004.<br />

113. Valliant, J. F. et al., The medicinal chemistry of carboranes, Coord. Chem. Rev., 232, 173, 2002.<br />

114. Hawthorne, M. F., The role of chemistry in the development of cancer therapy <strong>by</strong> the boron-neutron<br />

capture reaction, Angew. Chem. Int. Ed. Engl., 105, 997, 1993.<br />

115. Moraes, A. M., Santanaand, M. H. A., and Carbonell, R. G., Preparation and characterization of<br />

liposomal systems entrapping the boronated compound o-carboranylpropylamine,<br />

J. Microencapsul., 16, 647, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 101<br />

116. Moraes, A. M., Santana, M. H. A., and Carbonell, R. G., Characterization of liposomal systems<br />

entrapping boron-containing compounds in response to pH gradients, In Biofunctional Membranes,<br />

Proceedings of the International Conference on Biofunctional Membranes, Butterfield, A., Ed.,<br />

Plenum Press, New York, p.259, 1996.<br />

117. Shelly, K. et al., Model studies directed toward the boron neutron-capture therapy of cancer: boron<br />

delivery to murine tumors with liposomes, Proc. Natl. Acad. Sci. USA, 89, 9039, 1992.<br />

118. Feakes, D. A. et al., [Na 3B 20H 17NH 3]: Synthesis and liposomal delivery to murine tumors, Proc.<br />

Natl. Acad. Sci. USA, 91, 3029, 1994.<br />

119. Hawthorne, M. F., Shelly, K., and Li, F., The versatile chemistry of the B 20H 18] 2K ions: novel<br />

reactions and structural motifs, Chem. Commun. (Camb.), 547, 2002.<br />

120. Feakes, D. A. et al., Synthesis and in vivo murine evaluation of [Na 41-(1 0 -B 10H 9)-6-SHB 10H 8]asa<br />

potential agent for boron neutron capture therapy, Proc. Natl Acad. Sci. USA, 96, 6406, 1999.<br />

121. Watson-Clark, R. A. et al., Model studies directed toward the application of boron neutron capture<br />

therapy to rheumatoid arthritis: boron delivery <strong>by</strong> liposomes in rat collagen-induced arthritis, Proc.<br />

Natl. Acad. Sci. USA, 95, 2531, 1998.<br />

122. Feakes, D. A., Spinler, J. K., and Harris, F. R., Synthesis of boron-containing cholesterol derivatives<br />

for incorporation into unilamellar liposomes and evaluation as potential agents for BNCT,<br />

Tetrahedron, 55, 11177, 1999.<br />

123. Thirumamagal, B. T. S., Zhao, X. B., Bandyopadhyaya, A. K., Narayanasamy, S., Johnsamuel, J.,<br />

Tiwari, R., Golightly, D. W., Patel, V., Jehning, B. T., Backer, M. V., Barth, R. F., Lee, R. J., Backer,<br />

J. M., and Tjarks, W., Receptor-targeted liposomal delivery of boron-containing cholesterol mimics<br />

for boron neutron capture therapy (BNCT), Bioconjug. Chem., <strong>2006</strong>, in press.<br />

124. Pan, G., Oie, S., and Lu, D. R., Uptake of the carborane derivative of cholesteryl ester <strong>by</strong> glioma<br />

cancer cells is mediated through LDL receptors, Pharm. Res., 21, 1257, 2004.<br />

125. Maletinska, L. et al., Human glioblastoma cell lines: levels of low-density lipoprotein receptor and<br />

low-density lipoprotein receptor-related protein, Cancer Res., 60, 2300, 2000.<br />

126. Nygren, C. et al., Increased levels of cholesterol esters in glioma tissue and surrounding areas of<br />

human brain, Br. J. Neurosurg., 11, 216, 1997.<br />

127. Leppala, J. et al., Accumulation of 99m Tc-low-density lipoprotein in human malignant glioma, Br.<br />

J. Cancer, 71, 383, 1995.<br />

128. Peacock, G. et al., In vitro uptake of a new cholesteryl carborane ester compound <strong>by</strong> human glioma<br />

cell lines, J. Pharm. Sci., 93, 13, 2004.<br />

129. Wei, Q., Kullberg, E. B., and Gedda, L., Trastuzumab-conjugated boron-containing liposomes for<br />

tumor-cell targeting; development and cellular studies, Int. J. Oncol., 23, 1159, 2003.<br />

130. Bohl Kullberg, E. et al., Development of EGF-conjugated liposomes for targeted delivery of boronated<br />

DNA-binding agents, Bioconjug. Chem., 13, 737, 2002.<br />

131. Stephenson, S. M. et al., Folate receptor-targeted liposomes as possible delivery vehicles for boron<br />

neutron capture therapy, Anticancer Res., 23, 3341, 2003.<br />

132. Yanagie, H. et al., Boron neutron capture therapy using 10 B entrapped anti-CEA immunoliposome,<br />

Hum. Cell, 2, 290, 1989.<br />

133. Yanagie, H. et al., Application of boronated anti-CEA immunoliposome to tumour cell growth<br />

inhibition in in vitro boron neutron capture therapy model, Br. J. Cancer, 63, 522, 1991.<br />

134. Yanagie, H. et al., Inhibition of growth of human breast cancer cells in culture <strong>by</strong> neutron capture<br />

using liposomes containing 10 B, Biomed. Pharmacother., 56, 93, 2002.<br />

135. Xu, L., Boron neutron capture therapy of human gastric cancer <strong>by</strong> boron-containing immunoliposomes<br />

under thermal neutron irradiation (in Chinese), Zhonghua Yi Xue Za Zhi., 71, 568, 1991.<br />

136. Xu, L., Zhang, X. Y., and Zhang, S. Y., In vitro and in vivo targeting therapy of immunoliposomes<br />

against human gastric cancer (in Chinese), Zhonghua Yi Xue Za Zhi., 74, 83, 1994.<br />

137. Pan, X. Q., Wang, H., and Lee, R. J., Boron delivery to a murine lung carcinoma using folate<br />

receptor-targeted liposomes, Anticancer Res., 22, 1629, 2002.<br />

138. Pan, X. Q. et al., Boron-containing folate receptor-targeted liposomes as potential delivery agents for<br />

neutron capture therapy, Bioconjug. Chem., 13, 435, 2002.<br />

139. Sudimack, J. J. et al., Folate receptor-mediated liposomal delivery of a lipophilic boron agent to<br />

tumor cells in vitro for neutron capture therapy, Pharm. Res., 19, 1502, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


102<br />

Nanotechnology for Cancer Therapy<br />

140. Bohl Kullberg, E. et al., Introductory experiments on ligand liposomes as delivery agents for boron<br />

neutron capture therapy, Int. J. Oncol., 23, 461, 2003.<br />

141. Kullberg, E. B., Nestor, M., and Gedda, L., Tumor-cell targeted epiderimal growth factor liposomes<br />

loaded with boronated acridine: uptake and processing, Pharm. Res., 20, 229, 2003.<br />

142. Mehvar, R., Dextrans for targeted and sustained delivery of therapeutic and imaging agents,<br />

J. Control. Release, 69, 1, 2000.<br />

143. Mehvar, R., Recent trends in the use of polysaccharides for improved delivery of therapeutic agents:<br />

pharmacokinetic and pharmacodynamic perspectives, Curr. Pharm. Biotechnol., 4, 283, 2003.<br />

144. Chau, Y., Tan, F. E., and Langer, R., Synthesis and characterization of dextran-peptide–methotrexate<br />

conjugates for tumor targeting via mediation <strong>by</strong> matrix metalloproteinase II and matrix metalloproteinase<br />

IX, Bioconjug. Chem., 15, 931, 2004.<br />

145. Zhang, X. and Mehvar, R., Dextran–methylprednisolone succinate as a prodrug of methylprednisolone:<br />

Plasma and tissue disposition, J. Pharm. Sci., 90, 2078, 2001.<br />

146. Larsson, B., Gabel, D., and Borner, H. G., Boron-loaded macromolecules in experimental physiology:<br />

tracing <strong>by</strong> neutron capture radiography, Phys. Med. Biol., 29, 361, 1984.<br />

147. Gabel, D. and Walczyna, R., B-Decachloro-o-carborane derivatives suitable for the preparation of<br />

boron-labeled biological macromolecules, Z. Naturforsch., C, 37, 1038, 1982.<br />

148. Pettersson, M. L. et al., In vitro immunological activity of a dextran-boronated monoclonal antibody,<br />

Strahlenther. Onkol., 165, 151, 1989.<br />

149. Ujeno, Y. et al., The enhancement of thermal-neutron induced cell death <strong>by</strong> 10-boron dextran,<br />

Strahlenther. Onkol., 165, 201, 1989.<br />

150. Pettersson, M. L. et al., Immunoreactivity of boronated antibodies, J. Immunol. Methods, 126, 95,<br />

1990.<br />

151. Holmberg, A. and Meurling, L., Preparation of sulfhydrylborane-dextran conjugates for boron<br />

neutron capture therapy, Bioconjug. Chem., 4, 570, 1993.<br />

152. Carlsson, J. et al., Strategy for boron neutron capture therapy against tumor cells with overexpression<br />

of the epidermal growth factor-receptor, Int. J. Radiat. Oncol. Biol. Phys., 30, 105, 1994.<br />

153. Gedda, L. et al., Development and in vitro studies of epidermal growth factor-dextran conjugates for<br />

boron neutron capture therapy, Bioconjug. Chem., 7, 584, 1996.<br />

154. Mehta, S. C. and Lu, D. R., Targeted drug delivery for boron neutron capture therapy, Pharm. Res.,<br />

13, 344, 1996.<br />

155. Olsson, P. et al., Uptake of a boronated epidermal growth factor-dextran conjugate in CHO xenografts<br />

with and without human EGF-receptor expression, Anti-Cancer Drug Des., 13, 279, 1998.<br />

156. Novick, S. et al., Linkage of boronated polylysine to glycoside moieties of polyclonal antibody;<br />

boronated antibodies as potential delivery agents for neutron capture therapy, Nucl. Med. Biol., 29,<br />

159, 2002.<br />

157. Ferro, V. A., Morris, J. H., and Stimson, W. H., A novel method for boronating antibodies without<br />

loss of immunoreactivity, for use in neutron capture therapy, Drug Des. Discov., 13, 13, 1995.<br />

158. Sano, T., Boron-enriched streptavidin potentially useful as a component of boron carriers for neutron<br />

capture therapy of cancer, Bioconjug. Chem., 10, 905, 1999.<br />

159. Thomas, J. and Hawthorne, M. F., Dodeca(carboranyl)-substituted closomers: Toward unimolecular<br />

nanoparticles as delivery vehicles for BNCT, Chem. Commun. (Camb.), 1884, 2001.<br />

160. Pardridge, W. M., Vector-mediated drug delivery to the brain, Adv. Drug Deliv. Rev., 36, 299, 1999.<br />

161. Pardridge, W. M., Blood–brain barrier biology and methodology, J. Neurovirol., 5, 556, 1999.<br />

162. Hatakeyama, H. et al., Factors governing the in vivo tissue uptake of transferrin-coupled polyethylene<br />

glycol liposomes in vivo, Int. J. Pharm., 281, 25, 2004.<br />

163. Yanagie, H. et al., Accumulation of boron compounds to tumor with polyethylene–glycol binding<br />

liposome <strong>by</strong> using neutron capture autoradiography, Appl. Radiat. Isot., 61, 639, 2004.<br />

164. Maruyama, K. et al., Targetability of novel immunoliposomes modified with amphipathic poly<br />

(ethylene glycol)s conjugated at their distal terminals to monoclonal antibodies, Biochim.<br />

Biophys. Acta, 1234, 74, 1995.<br />

165. Yang, W. et al., Evaluation of systemically administered radiolabeled epidermal growth factor as a<br />

brain tumor targeting agent, J. Neurooncol., 55, 19, 2001.<br />

166. Bobo, R. H. et al., Convection-enhanced delivery of macromolecules in the brain, Proc. Natl Acad.<br />

Sci. USA, 91, 2076, 1994.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Neutron Capture Therapy of Cancer 103<br />

167. Groothuis, D. R., The blood–brain and blood–tumor barriers: A review of strategies for increasing<br />

drug delivery, Neurooncology, 2, 45, 2000.<br />

168. Vogelbaum, M. A., Convection enhanced delivery for the treatment of malignant gliomas: symposium<br />

review, J. Neurooncol., 73, 57, 2005.<br />

169. Husain, S. R. and Puri, R. K., Interleukin-13 receptor-directed cytotoxin for malignant glioma<br />

therapy: from bench to bedside, J. Neurooncol., 65, 37, 2003.<br />

170. Kunwar, S., Convection enhanced delivery of IL13-PE38QQR for treatment of recurrent malignant<br />

glioma: presentation of interim findings from ongoing phase 1 studies, Acta Neurochir. Suppl., 88,<br />

105, 2003.<br />

171. Wikstrand, C. J. et al., Monoclonal antibodies against EGFRvIII are tumor specific and react with<br />

breast and lung carcinomas and malignant gliomas, Cancer Res., 55, 3140, 1995.<br />

172. Wikstrand, C. J. et al., Cell surface localization and density of the tumor-associated variant of the<br />

epidermal growth factor receptor, EGFRvIII, Cancer Res., 57, 4130, 1997.<br />

173. Barth, R. F., Wu, G., Kawabata, S., Sferra, T. J., Bandyopadhyaya, A. K., Tjarks, W., Ferketich,<br />

A. K., Binns, P. J., Riley, K. J., Coderre, J. A., Ciesielski, M. J., Fenstermaker, R. A., Wikstrand, C. J.<br />

et al., Molecular targeting and treatment of EGFRvIII positive gliomas using boronated monoclonal<br />

antibody L8A4, Clin. Cancer Res., 12, 3792, <strong>2006</strong>.<br />

174. Cokgor, I. et al., Phase I trial results of iodine-131-labeled antitenascin monoclonal antibody 81C6<br />

treatment of patients with newly diagnosed malignant gliomas, J. Clin. Oncol., 18, 3862, 2000.<br />

175. Akabani, G. et al., Dosimetry and radiographic analysis of 131I-labeled anti-tenascin 81C6 murine<br />

monoclonal antibody in newly diagnosed patients with malignant gliomas: a phase II study, J. Nucl.<br />

Med., 46, 1042, 2005.<br />

176. Laske, D. W., Youle, R. J., and Oldfield, E. H., Tumor regression with regional distribution of the<br />

targeted toxin TF-CRM107 in patients with malignant brain tumors, Nat. Med., 3, 1362, 1997.<br />

177. Sampson, J. H. et al., Progress report of a Phase I study of the intracerebral microinfusion of a<br />

recombinant chimeric protein composed of transforming growth factor (TGF)-alpha and a mutated<br />

form of the Pseudomonas exotoxin termed PE-38 (TP-38) for the treatment of malignant brain<br />

tumors, J. Neurooncol., 65, 27, 2003.<br />

178. Weber, F. et al., Safety, tolerability, and tumor response of IL4-Pseudomonas exotoxin (NBI-3001)<br />

in patients with recurrent malignant glioma, J. Neurooncol., 64, 125, 2003.<br />

179. Weber, F. W. et al., Local convection enhanced delivery of IL4-Pseudomonas exotoxin (NBI-3001)<br />

for treatment of patients with recurrent malignant glioma, Acta Neurochir. Suppl., 88, 93, 2003.<br />

180. Ferrari, M., Nanovector therapeutics, Curr. Opin. Chem. Biol., 9, 343, 2005.<br />

181. Ferrari, M., Cancer nanotechnology: opportunities and challenges, Nat. Rev. Cancer, 5, 161, 2005.<br />

182. Blue, T. E. and Yanch, J. C., Accelerator-based epithermal neutron sources for boron neutrom<br />

capture therapy of brain tumors, J. Neurooncol., 62, 19, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


7<br />

CONTENTS<br />

Preclinical Characterization of<br />

Engineered Nanoparticles<br />

Intended for Cancer<br />

Therapeutics<br />

Anil K. Patri, Marina A. Dobrovolskaia, Stephan T. Stern, and<br />

Scott E. McNeil<br />

7.1 Introduction ......................................................................................................................... 106<br />

7.1.1 Physicochemical Characterization .......................................................................... 106<br />

7.1.2 In Vitro Characterization ........................................................................................ 107<br />

7.1.3 In Vivo Characterization ......................................................................................... 107<br />

7.2 Physicochemical Characterization ...................................................................................... 107<br />

7.2.1 Physicochemical Characterization Strategies ......................................................... 108<br />

7.2.2 Instrumentation........................................................................................................ 108<br />

7.2.2.1 Spectroscopy ............................................................................................ 108<br />

7.2.2.2 Chromatography ....................................................................................... 109<br />

7.2.2.3 Microscopy ............................................................................................... 109<br />

7.2.2.4 Size and Size Distribution ....................................................................... 110<br />

7.2.2.5 Surface Characteristics............................................................................. 112<br />

7.2.2.6 Functionality............................................................................................. 112<br />

7.2.2.7 Composition and Purity ........................................................................... 113<br />

7.2.2.8 Stability .................................................................................................... 114<br />

7.3 In Vitro Pharmacological and Toxicological Assessments................................................ 115<br />

7.3.1 Special Considerations ............................................................................................ 115<br />

7.3.2 In Vitro Target-Organ Toxicity .............................................................................. 116<br />

7.3.3 Cytotoxicity ............................................................................................................. 117<br />

7.3.4 Oxidative Stress....................................................................................................... 118<br />

7.3.5 Apoptosis and Mitochondrial Dysfunction ............................................................. 119<br />

7.3.6 Proteomics and Toxicogenomics ............................................................................ 121<br />

7.4 In Vivo Pharmacokinetic and Toxicological Assessments ................................................ 121<br />

7.5 Immunotoxicity ................................................................................................................... 124<br />

This project has been funded in whole or in part with federal funds from the National Cancer Institute, National Institutes of<br />

Health, under contract N01-CO-12400. The content of this publication does not necessarily reflect the views or policies of the<br />

Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply<br />

endorsement <strong>by</strong> the U.S. Government.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

105


106<br />

7.5.1 Applicability of Standard Immunological Methods for Nanoparticle Evaluation<br />

and Challenges Specific to Nanoparticle Characterization .................................... 125<br />

7.5.1.1 Blood Contact Properties ......................................................................... 125<br />

7.5.2 Immunogenicity....................................................................................................... 128<br />

7.6 Conclusions ......................................................................................................................... 128<br />

References .................................................................................................................................... 129<br />

7.1 INTRODUCTION<br />

As discussed in other chapters of this book, nanotechnology offers the research community the<br />

potential to significantly transform cancer diagnostics and therapeutics. Our ability to manipulate<br />

the biological and physicochemical properties of nanomaterial allows for more efficient drug<br />

targeting and delivery, resulting in greater potency and specificity, and decreased adverse side<br />

effects. The combinatorial possibilities of these multifunctional platforms have been the focus of<br />

considerable research and funding, but realization of their use in clinical trials is highly dependent<br />

on rigorous preclinical characterization to meet regulatory provisions and elucidated structure–<br />

activity relationships (SARs). A rational characterization strategy for biomedical nanoparticles<br />

contains three essential components (see Figure 7.1): physicochemical characterization, in vitro<br />

assays, and in vivo studies.<br />

7.1.1 PHYSICOCHEMICAL CHARACTERIZATION<br />

One of the major criticisms of early biomedical nanotechnology research was the general lack of<br />

physicochemical characterization that did not allow for the meaningful interpretation of resulting<br />

data or inter-laboratory comparisons. Traditional small molecule drugs are characterized <strong>by</strong> data that<br />

contribute to the chemistry, manufacturing and controls (CMC) section of the investigational new<br />

drug (IND) application with Food and Drug Administration (FDA), which include their molecular<br />

weight, chemical composition, identity, purity, solubility, and stability. The instrumentation to<br />

ascertain these properties has been well established, and the techniques are standardized. Engineered<br />

nanomaterials have dimensions between small molecules and bulk materials and often exhibit<br />

different physical and chemical properties than their counterparts. These physical and chemical<br />

Physical<br />

Characterization:<br />

Size<br />

Shape<br />

Composition<br />

Molecular weight<br />

Surface chemistry<br />

Identity<br />

Purity<br />

Stability<br />

Solubility<br />

In Vitro:<br />

Pharmacology<br />

Blood contact properties<br />

Effects on immune cell<br />

function<br />

Cytotoxicity<br />

Mechanistic toxicology<br />

Sterility<br />

In Vivo:<br />

ADME<br />

Safety<br />

Efficacy<br />

FIGURE 7.1 An assay cascade for preclinical characterization of nanomaterials.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 107<br />

properties influence the biological activity of nanoparticles and may depend on parameters such as<br />

particle size, size distribution, surface area, surface charge, surface functionality, shape, and aggregation<br />

state. Additionally, because most nanoparticle concepts are multifunctional, the<br />

distribution of targeting, imaging, and therapeutic components can also have dramatic effects on<br />

nanoparticle biological activity. There is a need to establish and standardize techniques to define<br />

these nanoparticle attributes.<br />

There is now ample evidence that size and surface characteristics can dramatically affect<br />

nanoparticle behavior in biological systems. 1–4 For instance, a decrease in particle size leads to<br />

an exponential increase in surface area per unit mass, and an attendant increase in the availability of<br />

reactive groups on the surface. Nanoparticles with cationic surface character have a notably<br />

increased ability to cross the blood–brain barrier compared to nanoparticles with anionic surfaces. 5<br />

In general, surface area, rather than mass, provides a better fit of dose–response relationships in<br />

toxicity studies for particles of various sizes. 6,7 Physicochemical characterization of properties,<br />

such as size, surface area, surface chemistry, and aggregation/agglomeration state, can provide the<br />

basis for better understanding of SARs.<br />

7.1.2 IN VITRO CHARACTERIZATION<br />

In vitro assays enable the isolation and analysis of specific biological and mechanistic pathways<br />

under controlled conditions. While many in vitro assays for nanomaterials will be similar to those<br />

used for traditional drugs, others will address mechanisms more specific to nanoparticles, such as<br />

oxidative stress. Noncellular assays measuring processes, such as protein adsorption, will also be an<br />

important accompaniment to cell-based assays. For example, monitoring the profile of serum<br />

proteins that absorb to nanoparticles in an in vitro environment may further our understanding of<br />

how nanoparticles interact with components of the reticuloendothelial system (RES) in vivo. 8,9<br />

Additionally, proteomics and toxicogenomics can be employed to potentially identify biomarkers<br />

of toxicity related to nanomaterial exposure. 10<br />

7.1.3 IN VIVO CHARACTERIZATION<br />

In vivo studies must be conducted to better understand the safety and behavior of nanoparticles in a<br />

living organism. As with any new chemical entity (NCE), the nanoparticle formulations’ pharmacological<br />

and toxicological properties (i.e., ADME/Tox) need to be thoroughly characterized.<br />

In vivo studies should include examination of nanoparticles’ effects on various organs and<br />

systems, such as the liver, heart, kidney, and immune system.<br />

In this chapter, we outline the scientific rationale underlying the development of an assay<br />

cascade, with special attention paid to the selection and adaptations of assays and analytical protocols<br />

needed to extract meaningful efficacy and safety data from nanomaterials. These are presented in the<br />

following four sections: (1) physicochemical characterization, (2) in vitro pharmacology and toxicology<br />

assessment, (3) in vivo pharmacology and toxicology assessment, and (4)<br />

immunotoxicity. Standardized characterization of nanomaterials will facilitate better inter-laboratory<br />

comparison of results and will enhance the quality of scientific data submitted <strong>by</strong> investigators in<br />

support of their IND applications.<br />

7.2 PHYSICOCHEMICAL CHARACTERIZATION<br />

The physicochemical characteristics of nanomaterials can affect their cellular uptake, binding to<br />

blood proteins, access to target sites, and ability to cause damage to cells and tissue. 11 Standard<br />

methods for physicochemical characterization of nanomaterials will provide the basis for rational<br />

product development as well as consistent and interpretable results for tests of efficacy and safety.<br />

Few examples of standard characterization criteria exist in the literature, and there is as yet no<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


108<br />

consensus as to what measurement criteria are appropriate for any given nanomaterial product.<br />

However, it is clear that the diversity and complexity of nanomaterials used in biomedical applications<br />

dictates a more comprehensive and strategic approach to characterization than has been<br />

applied to date.<br />

There are many varieties of nanomaterials currently being investigated for biomedical applications,<br />

especially for cancer diagnosis, imaging, and targeted drug delivery. These nanomaterials<br />

may be classified under several broad categories:<br />

† Organic nanoparticles (e.g., dendrimers, polymers, functionalized fullerenes)<br />

† Inorganic nanoparticles and organic–inorganic hybrids (e.g., iron oxide core particles,<br />

quantum dots)<br />

† Liposomes and other biological nanomaterials<br />

Each category of nanomaterial has a distinctly different composition that gives rise to different<br />

physical properties, such as solubility, stability, surface characteristics, and functional capabilities.<br />

Even within a single category, there can be a tremendous variety of product compositions, each<br />

with unique physical and chemical characteristics, and each requiring a different strategy for<br />

measuring those characteristics. This section examines the various categories of nanoparticles,<br />

and the tools and instrumentation available to address physicochemical characterization.<br />

7.2.1 PHYSICOCHEMICAL CHARACTERIZATION STRATEGIES<br />

Successful characterization strategies will enable one to begin associating the physicochemical<br />

properties of a nanomaterial with its in vivo behavior (i.e., SARs). This is an important step in the<br />

development of any material used for medical applications. For small molecules, the basis of most<br />

traditional drugs, the characterization techniques have been well established and standardized to<br />

determine their attributes, such as melting point, boiling point, molecular weight and structure,<br />

identity, composition, solubility, purity, and stability. These characteristics are measured and<br />

adequately defined using elemental analysis, mass spectrometry (MS), nuclear magnetic resonance<br />

(NMR), ultraviolet-visible (UV–vis) spectrophotometry, infrared (IR) spectroscopy, high-performance<br />

liquid chromatography (HPLC), gas chromatography (GC), capillary electrophoresis<br />

(CE), polarimetry, and other common analytical methods. Each of these individual techniques<br />

provides unique information about the sample, while together they provide the foundation for<br />

product quality control, manufacturing, and regulatory approval.<br />

Many of the techniques used to characterize small molecules apply to nanomaterials. However,<br />

due to the composite nature of nanomaterials, the definition and measurement of these attributes can<br />

be quite different. To fully understand the attributes of a nanomaterial, additional characterizations<br />

are needed, such as size, surface chemistry, surface area, polydispersity, and zeta potential (see<br />

Figure 7.2). A comprehensive analysis of these properties is necessary to better understand in vivo<br />

effects and to allow for greater consistency and reproducibility in their preparation. The requirements<br />

set <strong>by</strong> regulatory bodies for quality control and consistency of biomedical nanomaterials are<br />

likely to be as stringent as those for small molecule preparations, but the path to verifying quality<br />

will require a more sophisticated approach. At the core of this analysis is an array of tools and<br />

instrumentation that are particularly well suited to measuring the properties of nanomaterials.<br />

7.2.2 INSTRUMENTATION<br />

7.2.2.1 Spectroscopy<br />

Nanotechnology for Cancer Therapy<br />

Many traditional analytical methods can be applied to the characterization of nanomaterials. For<br />

example, NMR is extensively used to characterize dendrimers, polymers, and fullerenes<br />

derivatives, and provides unique information on the structure, purity, and functionality. 12–14<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 109<br />

O O<br />

N<br />

H OH<br />

HO<br />

O<br />

O<br />

In addition, the average number of terminal capping groups, number of small molecule ligands, and<br />

drugs in a multifunctional nanomaterial can be ascertained <strong>by</strong> comparing the integration values<br />

with chemical shifts unique to the ligands. UV–vis absorption spectrophotometry is also extensively<br />

used to identify and quantify the chromophore present in the preparation <strong>by</strong> using its<br />

extinction coefficient. Spectrofluorimetry is used in cases where the material has inherent fluorescence<br />

(such as quantum dots) or labeled with a fluorescence probe. Matrix-assisted laser<br />

desorption ionization time-of-flight (MALDI-TOF) MS is used extensively for macromolecules,<br />

dendrimers, and polymers to determine the molecular weight and utilize novel matrices to minimize<br />

the fragmentation of the macromolecule before reaching the detector. In the case of lower generation<br />

dendrimers, the presence of impurities, incomplete reaction, and reaction <strong>by</strong>products can be<br />

easily determined using MS.<br />

7.2.2.2 Chromatography<br />

O<br />

Small molecules<br />

• Elemental analysis<br />

• Mass spectrometry<br />

• NMR spectroscopy<br />

• UV−vis spectroscopy<br />

• IR spectroscopy<br />

• HPLC<br />

• GC<br />

• Polarimetry<br />

O<br />

Liquid chromatography methods such as analytical HPLC and size-exclusion chromatography<br />

(SEC; also called gel-permeation chromatography or GPC) utilize a column to separate components<br />

of a mixture in a liquid mobile phase based on their interaction with a solid stationary phase. The<br />

eluents are passed through UV–vis and fluorescence detectors with a flow cell where the absorbance<br />

and fluorescence is recorded to determine the purity of the sample. Although these techniques are<br />

suitable for stable polymers, dendrimers, 15 functionalized fullerenes, and protein- and peptide-based<br />

nanomaterial, they are not suitable for particles that degrade under experimental conditions or have<br />

excessive nonspecific binding to the solid matrix.<br />

7.2.2.3 Microscopy<br />

O OH<br />

H<br />

O<br />

O<br />

–H<br />

O<br />

Physicochemical Parameters<br />

• Composition<br />

• Physical properties<br />

• Chemical properties<br />

• Identification<br />

• Quality<br />

• Purity<br />

• Stability<br />

Liposome<br />

Dendrimer Gold nanoshell<br />

Quantum dot Fullerene<br />

Nanomaterial<br />

• Microscopy (AFM, TEM, SEM)<br />

• Scattering (DLS, MALLS, SAXS,<br />

SANS), FCS, AU<br />

• SEC, FFF<br />

• Electrophoresis (CE, PAGE)<br />

• Zeta potential<br />

• Fluorimetry<br />

FIGURE 7.2 Physicochemical characterization methods and instrumentation for small molecules and<br />

nanotechnology.<br />

Scanning probe microscopy (SPM) techniques can be employed to measure the size, topography,<br />

composition, and structural properties of nanoparticles. Related techniques such as scanning<br />

tunneling microscopy (STM), electric field gradient microscopy (EFM), scanning thermal<br />

microscopy, and magnetic field microscopy (MFM) combined with atomic force microscopy<br />

(AFM), can be used to investigate the structural, electronic, thermal, and magnetic properties of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


110<br />

a nanomaterial. AFM uses a nanoscale probe to detect the inter-atomic forces and interactions<br />

between the probe and the material being analyzed and is capable of determining size and shape<br />

within a spatial resolution of a few angstroms. 16 Apart from the ability to measure the particle size<br />

in a dry state as well as in aqueous and physiological conditions, AFM is a useful tool to probe the<br />

interaction of nanoparticles with supported lipid bilayers. This technique has been successfully<br />

used to compare nanoparticle interactions in in vitro cell assays. 17,18 The ability to image under<br />

physiological conditions makes AFM a powerful tool for the characterization of nanoparticles in a<br />

dynamic, biological context. A variant of this method, molecular recognition force microscopy<br />

(MRFM), can be employed to study the specific ligand–receptor interactions between nanoparticles<br />

and their biological targets.<br />

Optical microscopy techniques are useful at the micron scale and are extensively used for<br />

imaging structural features. Fluorescence and confocal microscopy may be used to determine<br />

cellular binding and internalization of fluorescent-labeled nanoparticles 19 or those that are inherently<br />

fluorescent, such as quantum dots. But a more precise analysis of nanomaterial size and other<br />

direct measurements of physical properties will require a more sophisticated and specialized set of<br />

microscopic and spectroscopic techniques.<br />

Scanning electron microscopy (SEM) provides information on the size, size distribution, shape,<br />

and density of nanomaterials. Transmission electron microscopy (TEM) and high-resolution TEM<br />

are more powerful than SEM in providing details at the atomic scale and can yield information<br />

regarding the crystal structure, quality, and grain size. TEM can be coupled with other characterization<br />

tools, such as electron energy loss spectrometry (EELS) or energy dispersive x-ray<br />

spectrometry (EDS), to provide additional information on the electronic structure and elemental<br />

composition of nanomaterials. Samples for TEM are evaluated dry or in a frozen state, under highvacuum<br />

conditions. Nanoparticles analyzed <strong>by</strong> this instrument must therefore be stable under these<br />

extreme conditions. Additionally, while considered a gold standard of microscopic characterization<br />

methods, TEM requires a great deal of skill and time to obtain good data. In principle, when<br />

establishing characterization protocols, TEM can be used to validate characterization methods<br />

that are easier to use on a routine basis. Further description of analytical technologies as they<br />

apply to the measurement of specific nanomaterial properties is provided in the following sections.<br />

7.2.2.4 Size and Size Distribution<br />

Nanotechnology for Cancer Therapy<br />

Size is one of the critical parameters that dictate the absorption, biodistribution, and route of<br />

elimination for biomedical nanomaterials. 20 Generally, nanoparticles with dimensions of less<br />

than 5–10 nm are rapidly cleared after systemic administration, while particles from 10 to 70 nm<br />

in diameter may penetrate capillary walls throughout the body. 21,22 Larger particles 70–200 nm<br />

often remain in circulation for extended times. 22,23 This general correlation of biodistribution and<br />

elimination with respect to size may vary greatly depending on nanoparticle surface characteristics.<br />

Specifically in cancer applications, size is an important factor in the accumulation of therapeutic<br />

nanomaterials in tumors, usually as a result of enhanced permeation and retention (EPR),<br />

caused <strong>by</strong> local defects in the vasculature and poor lymphatic drainage. 24 Particle size can be<br />

precisely tuned to take advantage of this phenomenon and passively target and deliver a therapeutic<br />

payload to tumors. 25–27<br />

Depending on the category of the nanomaterial, synthesis and scale-up can be problematic.<br />

Most biomedical nanomaterials for therapeutic and diagnostic applications are complex and<br />

involve some combination of molecular self-assembly, encapsulation, and/or the use of nanosized<br />

metal or polymer cores, surfactants and/or proteins to impart solubility and functionality.<br />

Due to inherent variability in the manufacturing process, one rarely achieves a monodisperse,<br />

homogeneous product. It is therefore important to ascertain the precise size, size distribution,<br />

and polydispersity index (PDI) of the material. There are several techniques available to assess<br />

these parameters, including electron microscopy, AFM, and light scattering. Light scattering<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 111<br />

techniques can measure overall size and polydispersity of the particles. TEM is powerful in<br />

ascertaining the homogeneity of nanoparticles with encapsulated metals and in determining core<br />

size. With knowledge of nanoparticle geometry and size, surface area can also be estimated.<br />

For biological applications, it is important to measure the physical characteristics of the nanomaterial<br />

in isotonic solution at physiological pH and temperature. The hydrodynamic size can be<br />

measured under these conditions using dynamic light scattering (DLS) (also known as photon<br />

correlation spectroscopy [PCS] and quasi elastic light scattering [QELS]) and analytical ultracentrifugation<br />

(AU). In a DLS experiment, the effects of Brownian motion (particle movement caused<br />

<strong>by</strong> random collisions in solution) provide information on particle size and size distribution. The<br />

sample is illuminated with a laser, and the intensity fluctuations in the scattered light are analyzed<br />

and related to the size of the suspended particles. This technique is useful in determining whether<br />

the nanomaterial is monodisperse in size distribution. These data are influenced <strong>by</strong> the viscosity and<br />

the temperature of the medium, since Brownian motion depends on these factors. The pH of the<br />

medium and salt concentration may also affect the degree of agglomeration in some samples. With<br />

DLS, sample preparation is easy, the measurement is quick, and data are reproducible on larger<br />

sample volumes compared to microscopy techniques; however, better standardization of<br />

procedures, conditions, and data analysis tools will be required. Static light scattering provides<br />

information on molar mass and root-mean-squared (rms) radius for fractionated or monodisperse<br />

samples. One limitation of light scattering instruments is the inability to measure the size when the<br />

nanoparticles absorbs in the wavelength of the laser being used. Small-angle x-ray scattering<br />

(SAXS) 28 and small-angle neutron scattering (SANS) 29 can be used to measure the size, shape<br />

and orientation of components. Due to their cost and infrastructure requirements, there is limited<br />

availability of these instruments. For fluorescent nanomaterials such as quantum dots, size can be<br />

measured using fluorescence correlation spectroscopy (FCS). 30<br />

The hydrodynamic size of nanoparticles can also be measured with AU, which is traditionally<br />

used to measure the size of proteins. 31 The instrument spins the protein sample solution under high<br />

vacuum at a controlled speed and temperature while recording concentration distribution at set<br />

times. Even though this technique is designed to measure the size of proteins in solution, it has<br />

potential applications in the measurement of the hydrodynamic size of nanoparticles samples that<br />

are stable under the experimental conditions. Fractionation using SEC separates stable polymers<br />

into individual components and helps in the determination of the PDI. In the case of unfractionated<br />

samples, batch mode measurement provides averaged quantities such as weight-averaged molar<br />

mass and z-average rms radius. This technique is especially useful when combined with a refractive<br />

index detector to obtain absolute molecular weight for very high molecular weight polymers where<br />

traditional MS methods fail.<br />

In cases where the separation and fractionation of nanomaterial is not possible using a<br />

column with a stationary phase, such as when the nanomaterial may interact with the column<br />

packing material and render it unstable, asymmetric-flow field flow fractionation (AFFF) is<br />

useful. 32 In AFFF, separation occurs when the sample passes through a narrow channel with<br />

a cross-flow through a porous semi-permeable membrane. The faster moving smaller particles<br />

rise to the top of the flow and come out first followed <strong>by</strong> larger particles that stay closer to the<br />

membrane and migrate more slowly. One advantage in this method is that there is no stationary<br />

phase in the separation: the sample injected comes out intact with little loss of material due to<br />

nonspecific binding. This feature is particularly useful for less stable nanoparticles such as<br />

liposomes, or for polymer- or protein-coated metal nanoparticles that would otherwise interfere<br />

with the performance of a traditional GPC column. The efficiency of separation for AFFF is not<br />

as good as with GPC, but there have been recent improvements in instrumentation that are<br />

closing the gap in performance. For both GPC and AFFF, the quantity and hydrodynamic size of<br />

the nanoparticles are detected in eluted peaks <strong>by</strong> measuring absorbance, refractive index, and<br />

light scattering.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


112<br />

In addition to size, the shape of a nanoparticle may affect its distribution and absorption in the<br />

body. Spherical, tubular, plate-like, or nano-porous materials of the same composition can vary<br />

significantly in their surface energy, biological activity, and access to different physiological<br />

structures, such as cell walls, capillary vessels, etc. Methods such as AFM, SEM, TEM, and<br />

STM can be used to determine the distribution of shape in a nanoparticle preparation.<br />

7.2.2.5 Surface Characteristics<br />

Surface characteristics contribute to the nanoparticle’s solubility, aggregation tendency, ability to<br />

traverse biological barriers (such as a cell wall), biocompatibility, and targeting ability. The<br />

nanoparticle surface is also responsible for interaction and binding with plasma proteins in vivo,<br />

which in turn may alter the nanoparticle’s distribution and pharmacokinetics. For multifunctional<br />

nanoparticles, modifying agents are often attached to the surface to bind to receptors in target<br />

tissues and organs. The presence of charged functionalities on the nanoparticle surface may<br />

increase nonspecific uptake, making the preparation less effective in targeting. It has been shown<br />

that dendrimer nanoparticles displaying positively charged amine groups on their surface can be<br />

significantly more hemolytic and cytotoxic than nanoparticles displaying negatively charged<br />

carboxylates. 20 The negatively charged nanoparticles were also cleared more slowly from the<br />

blood compared to positively charged species, following intravenous administration to rats. 20<br />

Another potential effect of surface charge is to alter a nanoparticle’s ability to penetrate the<br />

blood–brain barrier. Studies have shown that for emulsifying wax nanoparticles, anionic surfaces<br />

were superior to neutral or cationic surfaces for penetration of the blood–brain barrier. 33<br />

Surface characteristics can be tuned to improve receptor binding, reduce toxicity, or alter<br />

biodistribution. For example, when the above-mentioned dendrimers were acetylated to neutralize<br />

exposed surface charges, the toxic effects of the nanoparticles were also neutralized. 20,34 Surface<br />

properties can also lead to toxicity through interaction with molecular oxygen, leading to oxidative<br />

stress and inflammation. Electron capture at the surface of the nanoparticle results in the formation<br />

of the superoxide radical, which can set off a cascade of reactions (e.g., through Fenton reaction or<br />

disumation) to generate reactive oxygen species (ROS). ROS generation has been studied extensively<br />

for inhaled nanoparticles, 11,35 and has been observed in engineered nanoparticles such as<br />

fullerenes, single walled nanotubes (SWNTs), and quantum dots. 6,36–44 Studies have shown a direct<br />

correlation between nanoparticle surface area and ROS-generating capacity and inflammatory<br />

effects. 11<br />

The nature and integrity of nanomaterial surfaces must be established through analytical<br />

measurements to ensure product quality and account for surface-dependent effects on biodistribution<br />

and toxicity. Potentiometric titrations provide crucial information on the net charge of a<br />

nanoparticle, and include zeta potential analysis, which provides information on the net charge<br />

and distribution under physiological conditions. Polyacrylamide gel electrophoresis (PAGE)<br />

analysis of dendrimers and other nanopolymers yields information on the molecular weight and<br />

the polydispersity of nanoparticles (such as trailing generations in dendrimer populations) based on<br />

their migration through the gel under an electric field. PAGE is also a powerful tool in the<br />

qualitative analysis of bioconjugates of nanomaterials with DNA, oligonucleotides, antibodies,<br />

and other ligands. Further analysis of the surface charge distribution and polydispersity of nanomaterials<br />

can be conducted using CE. MS is also effective in ascertaining the number and<br />

distribution of charges, especially for smaller and purer nanoparticles with known<br />

molecular weight.<br />

7.2.2.6 Functionality<br />

Nanotechnology for Cancer Therapy<br />

Analysis of the functional components of nanomaterials, such as targeting, imaging, and<br />

therapeutic agents, is critical to understand the in vivo efficacy of the preparation. Characteristic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 113<br />

features of functional components include their quantity, distribution, orientation, and activity. For<br />

targeting agents, a key advantage of their use in nanoparticles is their ability to provide increased<br />

avidity to the target due to polyvalency. The level of polyvalency and activity of targeting agents<br />

can be monitored using surface plasmon resonance (SPR) to measure the rate constants for nanoparticle<br />

association and dissociation. During preclinical development, the affinity of nanomaterial<br />

preparations for their target molecule/receptor can be analyzed using SPR and compared to data<br />

obtained for binding to cellular receptors in culture. 45<br />

The average number of targeting agents per nanoparticle has to be optimized for both solubility<br />

and binding affinity. Affinity chromatography or SEC can be employed with some nanoparticles to<br />

separate nanoparticles with targeting agents from those without targeting agents. In nanoparticles<br />

containing antibodies 19,46 or proteins, quantification can be achieved using an enzyme-linked<br />

immunosorbent assay (ELISA) or bicinchoninic acid assay (BCA) if the inherent property of the<br />

nanoparticle itself does not interfere with the assay. In the case of dendrimers, NMR has been<br />

successfully applied to analyze the average number of targeting agents <strong>by</strong> comparing the integration<br />

values of the signals associated with the targeting agents to those belonging to the<br />

dendrimer. This is still an averaged technique that cannot distinguish the distribution of targeting<br />

agent density on a population of nanoparticles.<br />

For targeted drug delivery applications, it is obviously important for both the targeting and<br />

therapeutic agents to be on the same particle. If the therapeutic has UV–vis absorption, it can be<br />

quantified using UV–vis spectroscopy with the extinction coefficient of the drug. HPLC analysis is<br />

possible in some cases to evaluate the amount of the drug present in a known amount of material,<br />

after isolating the drug from the sample.<br />

7.2.2.7 Composition and Purity<br />

Biomedical nanomaterials can be comprised of a wide variety of substances, including polymers,<br />

metals and metal oxides, lipids and other organic compounds, and large biomolecules such as<br />

protein or DNA. In most cases, the nanomaterials combine two or more of these substances,<br />

such as in a core or shell of a particle, and in encapsulated or conjugated material. Analysis of<br />

chemical composition will be critical for confirming the purity and homogeneity of nanomaterial<br />

product preparations.<br />

Elemental analysis, such as CHN analysis, is most often used to ascertain the purity of small<br />

molecules. For nanomaterials, elemental analysis can be used to determine the composition and<br />

ratios of different elements present in the sample. For example, this technique can be used to<br />

determine the amount of linker present, if a unique element (such as sulfur) has been employed<br />

in the synthesis. In the case of core–shell metal nanoparticles, the ratio of core to shell material<br />

ratios can be determined.<br />

Atomic absorption (AA) and atomic emission (AE) spectroscopies can also be utilized to<br />

determine the composition of nanomaterials. For imaging applications using iron oxide nanoparticles<br />

or gadolinium (Gd)-based chelates, composition analysis is very important to quantify metals<br />

present in the preparation which influence imaging efficacy. Inductively coupled plasmon optical<br />

emission spectroscopy (ICP-OES) is very sensitive to determine the amount of Gd in such contrast<br />

agent conjugates. 47 Specific T1/T2 relaxivities of magnetic resonance contrast agents can, of<br />

course, be assessed under in vitro conditions in the actual MRI instrument.<br />

The purity of synthetic small molecules can be determined with a high degree of certainty since<br />

the analyte usually consists of a single component. With nanomaterials, purity must be determined<br />

in the context of multiple layered, conjugated, and encapsulated components. Purity analysis must<br />

account for the presence of solvents, free metals and chelates, unconjugated therapeutic or other<br />

agents, precursors, dimers, etc., that result in artifacts and side products of the preparation. 48<br />

Characterization of the inhomogeneity in ligand distribution is very important for efficacy as<br />

well as testing batch-to-batch reproducibility. 49 Proper methods and techniques to detect the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


114<br />

presence of all these entities are required to ensure the purity and quality of nanomaterial preparations<br />

and to further expand our understanding of SARs.<br />

7.2.2.8 Stability<br />

Nanotechnology for Cancer Therapy<br />

The ability of multifunctional nanoparticles to combine targeting, therapeutic, and imaging modalities<br />

is a key aspect of their versatility and anticipated clinical impact. 50–52 With such complex<br />

compositions, the stability of all the components in nanoparticles is essential to their biological<br />

function. Premature release of any of the components from the composite preparation may render it<br />

ineffective. For example, in a nanodelivery system containing a targeting agent and a drug, the<br />

nanoparticles with the drug cannot bind to the desired targeting site if the targeting agent is<br />

prematurely cleaved or released. If the drug is prematurely released, even if the nanoparticle<br />

reaches its target, there will no longer be a therapeutic benefit. 53 For this reason, it is important<br />

to determine the in vitro functional component stability under physiological conditions.<br />

For a nanomaterial providing targeted or timed-release drug delivery with an encapsulated<br />

drug, the release profile should be determined at different ionic strength, pH, and temperature<br />

conditions. Examples of such conditions include the stability at pH 7.4, in buffers such as phosphate<br />

buffered saline (PBS), and serum at 378C. There are many nanoparticle designs being pursued<br />

which incorporate the selective release of components triggered <strong>by</strong> an external stimulus after<br />

targeted delivery. If a therapeutic attached to a nanoparticle uses a cleavable linkage, the efficiency<br />

of release should be determined under the expected cleavage conditions. 54<br />

In cases where a metal complex is used (for example, a Gd chelate for enhanced MRI contrast),<br />

the stability constants for the encapsulation or complexation should be determined, since any<br />

release of free heavy metal will increase the in vivo toxicity of the preparation. 55 The potential<br />

in vivo application of quantum dots has raised some concerns that the CdSe core might be exposed<br />

<strong>by</strong> the breakdown of its protective polymer or inorganic shell, releasing the highly toxic heavy<br />

metal Cd 2C ions into the bloodstream. 56 The quantum dot shells have been designed to be protective,<br />

but their long-term stability (e.g., susceptibility to Cd leaching) has not been established.<br />

Studies conducted on primary hepatocytes in vitro suggest that CdSe core quantum dots may be<br />

acutely toxic under certain conditions. 57 Other studies suggest that under physiological conditions,<br />

appropriately coated quantum dots do not expose the host organism to toxic levels of the core<br />

material. 58–60 Apparently conflicting evidence as to the safety of quantum dots highlights the<br />

necessity of clearly and objectively establishing the stability of these nanoparticles under physiological<br />

conditions using standardized methodologies.<br />

It is also important to determine the stability of the nanoparticle under nonphysiological<br />

conditions to account for the effects of short-term and long-term storage, lyophilization, ultrafiltration,<br />

thermal exposure, pH variation, freeze–thawing, and exposure to light.<br />

In summary, adequate physicochemical characterization of nanomaterials should be included<br />

as an essential requirement for preclinical characterization. Just as molecular characterization<br />

forms the basis of dosing and toxicity studies for small molecule therapeutics and diagnostic<br />

compounds, physicochemical characterization provides the foundation for dosing and toxicity<br />

studies for nanomaterials intended for clinical applications. Standardized protocols are being established<br />

<strong>by</strong> Standards Developing Organizations, such as the International Standards Organization<br />

(ISO) and American Society for Testing and Materials (ASTM), for characterizing the many types<br />

of biomedical nanomaterials being developed today for human use. Additionally, standardized<br />

reference material (SRM) will enable analytical technologies to be calibrated and protocols to be<br />

tested for consistency and to facilitate inter-laboratory comparisons.<br />

To better control for the results of in vivo studies of nanomaterial absorption, distribution,<br />

metabolism, elimination, and toxicity, it will be necessary to examine the material in the same<br />

physicochemical state as would be found under physiological conditions. Particle-specific attributes<br />

that should be evaluated include surface characteristics, chemical composition, shape, size,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 115<br />

and ligand dispersity. Additional properties that are influenced <strong>by</strong> experimental conditions include<br />

solubility, stability, protein binding, and aggregation state. Knowing the exact physiological conditions<br />

in different tissues and organs and developing a means to either replicate those conditions or<br />

measure physicochemical properties in situ is a significant challenge. But continued studies in this<br />

area will provide further data to elucidate the linkages between physicochemical characteristics of<br />

nanomaterials and their biological effects (i.e., SARs).<br />

7.3 IN VITRO PHARMACOLOGICAL AND TOXICOLOGICAL ASSESSMENTS<br />

Prior to filing an IND or investigational device exemption (IDE) application with the FDA and<br />

subsequent clinical testing in humans, a new product must be adequately studied for efficacy and<br />

safety using animal models. The cost- and labor-intensiveness of these in vivo studies impel drug<br />

and device researchers to make use of predictive in vitro methodologies wherever technology<br />

permits. In vitro models can serve as an initial assessment of a nanomaterial’s efficacy and absorption,<br />

distribution, metabolism, elimination, and toxicity (ADME/Tox), allowing a more strategic<br />

approach to animal studies. Used iteratively with in vivo studies, the two approaches can inform<br />

each other and help narrow investigations of the physiological and biochemical pathways that<br />

contribute to ADME/Tox behavior.<br />

A variety of cell-based in vitro systems are available, including perfused organs, tissue slices,<br />

cell cultures based on a single cell line or combination of cell lines, and primary cell preparations<br />

freshly derived from organ and tissue sources. In vitro models allow examination of biochemical<br />

mechanisms under controlled conditions, including specific toxicological pathways that may occur<br />

in target organs and tissues. Examples of mechanistic toxicological endpoints assessed in vitro<br />

include inhibition of protein synthesis and microtubule injury. These mechanistic endpoints can<br />

provide information not only as to the potential mechanisms of cell death, but also can identify<br />

compounds that may cause chronic toxicities that often results from sublethal mechanisms that may<br />

not cause overt toxicity in cytotoxicity assays. Common mechanistic paradigms associated with<br />

nanoparticle toxicity include oxidative stress, apoptosis, and mitochondrial dysfunction. Due to the<br />

nanoparticle- and approach-specific nature of pharmacology studies, it is beyond the scope of this<br />

chapter to discuss pharmacological assay specifics. Where appropriate models exist, chemotherapeutic<br />

efficacy can be examined in vitro. In certain cases, targeting of chemotherapeutic agents may<br />

be demonstrated as well, using optimized treatment/wash-out schemes in cell lines expressing the<br />

targeted receptor. Though nanoparticle metabolism or enzyme induction has yet to be demonstrated,<br />

certain nanomaterials with attractive chemistries may be subjected to phase-I/II<br />

metabolism and induction studies using cell-based, microsomal, and/or recombinant<br />

enzyme systems.<br />

7.3.1 SPECIAL CONSIDERATIONS<br />

Many of the standard methods used to evaluate biocompatibility of new molecular and chemical<br />

entities are fully applicable to nanoparticles. However, existing test protocols may require further<br />

development and laboratory validation before they become available for routine testing. Careful<br />

attention must be paid to potential sources of interference with analytical endpoints that may lead to<br />

false-positive or false-negative results. Nanoparticle interference could result from: interference<br />

with assay spectral measurements; inhibition/enhancement of enzymatic reactions 61,62 ; and absorption<br />

of reagents to nanoparticle surfaces. In the event of nanoparticle interference, additional<br />

sample preparation steps or alternative methods may be required.<br />

When evaluating the results of in vitro assays, it is important to recognize that dose–response<br />

relationships will not always follow a classical linear pattern. These atypical dose–response<br />

relationships have previously been attributed to shifts between the different mechanisms underlying<br />

the measured response. 63 In the case of nanomaterials, it is also important to bear in mind that<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


116<br />

concentration-dependent changes in the physical state (e.g., aggregation state, degree of protein<br />

binding) may also result in apparent nonlinearity.<br />

Another key consideration when evaluating the results of nanoparticle research is the impact of<br />

dose metric (e.g., mass, particle number, surface area), sample preparation (e.g., sonication), and<br />

experimental conditions (e.g., exposure to light) on the interpretation of results. For example,<br />

surface area or particle number may be a more appropriate metric than mass when comparing<br />

data generated for different sized particles. This has been shown to be the case for 20- and 250-nm<br />

titanium dioxide nanoparticles, in which lung inflammation in rats, as assessed <strong>by</strong> percentage of<br />

neutrophils in lung lavage fluid, correlated with total surface area rather than mass. 64 The importance<br />

of experimental conditions in study design is highlighted <strong>by</strong> an investigation of functionalized<br />

fullerenes, demonstrating that the cytotoxicity of dendritic and malonic acid functionalized fullerenes<br />

to human T-lymphocytes in vitro is enhanced <strong>by</strong> photoexcitation. 41 The standardization of<br />

these experimental variables should limit inter-laboratory variability and make data generated<br />

more comparable.<br />

7.3.2 IN VITRO TARGET-ORGAN TOXICITY<br />

Nanotechnology for Cancer Therapy<br />

A recently published report from the International Life Sciences Institute Research Foundation/Risk<br />

Science Institute Nanomaterial Toxicity Screening Working <strong>Group</strong> 73 recommends the inclusion<br />

of several specific in vitro assays in a standard protocol of safety assessment. Much of the report<br />

focused on toxicity screening for environmental exposure to nanoparticles and thus emphasized<br />

environmentally relevant exposure routes. However, in addition to the in vitro examination of<br />

so-called portal-of-entry tissues, the report expressed the need for inclusion of potential target<br />

organs. The liver and kidney were selected as ideal candidates for these initial in vitro target<br />

organ toxicity studies, since preliminary investigations (discussed below) have identified these as<br />

the primary organs involved in the accumulation, processing, and eventual clearance<br />

of nanoparticles.<br />

The liver has been identified in many studies as the primary organ responsible for<br />

reticuloendothelial capture of nanoparticles, often due to phagocytosis <strong>by</strong> Kupffer cells. 65–67 Fluorescein<br />

isothiocyanate-labeled polystyrene nanoparticles and radiolabeled dendrimers, for example,<br />

are rapidly cleared from the systemic circulation <strong>by</strong> hepatic uptake following intravenous injection.<br />

4,68 Hepatic uptake has also been shown to be a primary mechanism of hepatic clearance for<br />

parenterally administered fullerenes, dendrimers, and quantum dots. 20,69,70 In addition to hepatic<br />

accumulation, nanoparticles have also been shown to have a detrimental effect on liver function ex<br />

vivo and alter hepatic morphology. Hepatocytes isolated from rats intravenously administered<br />

polyalkylcyanoacrylate nanoparticles had diminished secretion of albumin and decreased<br />

glucose production. 71 This alteration in albumin synthesis was also observed in freshly isolated<br />

rat hepatocytes exposed to polyalkylcyanoacrylate nanoparticles. Dendrimers have also been<br />

shown to cause liver injury. Repeated dosing of mice with polyamidoamine (PAMAM) dendrimers<br />

resulted in vacuolization of the hepatic parenchyma, suggesting lysosomal dysfunction. 72<br />

Sprague–Dawley rat hepatic primary cells and human hepatoma Hep-G2, selected for in vitro<br />

hepatic target organ toxicity assays, have a long history of use in toxicological evaluation. 73–75<br />

Hep-G2 cells were chosen since they are a readily available hepatocyte cell line with high metabolic<br />

activity. 76 Rat hepatic primary cells were also chosen for toxicological studies, since hepatic<br />

primary cells in culture are more reflective of in vivo hepatocytes with regard to enzyme expression<br />

and specialized functions. One survey found rat hepatic primary cells up to ten times more sensitive<br />

to model hepatotoxic agents than established hepatic cell lines. 77 Rat hepatic primaries represent a<br />

suitable alternative to human hepatic primary cells, which are scarce, costly, and suffer from<br />

interindividual variability.<br />

Preliminary pharmacokinetic studies of parenterally administered, radiolabeled carbon nanotubes,<br />

dendritic fullerenes, and low generation dendrimers in rodents have identified urinary<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 117<br />

excretion as the principal mechanism of clearance. 78–80 A variety of engineered nanoparticles,<br />

including actinomycin D-loaded isobutylcyanoacrylate, doxorubicin-loaded cyanoacrylate, and<br />

dendrimer nanoparticles, have also been shown to distribute to renal tissue following parenteral<br />

administration in rodents. 68,81,82 In the case of the doxorubicin-loaded cyanoacrylate nanoparticles,<br />

doxorubicin renal distribution was increased due to capture of the nanoparticles <strong>by</strong> glomerular<br />

mesangial cells. This resulted in a shift in the primary target organ from the heart to the kidney.<br />

Doxorubicin-induced renal injury presented as a severe proteinuria. Kidney injury has been demonstrated<br />

for other nanomaterials as well. Nano-zinc particles, for example, caused severe histological<br />

alterations in murine kidneys and Q-dots were shown to be cytotoxic to African green monkey<br />

kidney cells. 83,84<br />

The porcine renal proximal tubule cell line, <strong>LLC</strong>-PK1, was selected as a representative kidney<br />

cell line, since it is readily available through ATCC and has been used extensively in nephrotoxicity<br />

screening and mechanistic studies. 85 The SD rat hepatic primary, <strong>LLC</strong>-PK1 and Hep-G2 cells are<br />

adherent, which can simplify sample preparation, and can be propagated in a 96-well plate format<br />

suitable for high-throughput screening. The 96-well format allows for detailed concentration–<br />

response curves and multiple controls to be run on the same microplate. These cell lines were<br />

subjected to a variety of in vitro assays, described below, to evaluate cytotoxicity, mechanistic<br />

toxicology, and pharmacology. These assays have been selected primarily for their superior performance,<br />

convenience, and adaptability in evaluating this new class of biomedical agent.<br />

7.3.3 CYTOTOXICITY<br />

Cell viability of adherent cell lines can be assessed <strong>by</strong> a variety of methods. 86 These methods fall<br />

broadly into four categories, assays that measure: (1) loss of membrane integrity; (2) loss of<br />

metabolic activity; (3) loss of monolayer adherence; and (4) cell cycle analysis. Data generated<br />

using these various viability assays can be used to identify cell lines susceptible to nanoparticle<br />

toxicity and potentially give clues as to the type (i.e., cytostatic/cytotoxic) and location of cellular<br />

injury. Many of the cytotoxicity assays discussed below are available as commercial kits. These kits<br />

should be used whenever feasible since they provide an extra level of quality control.<br />

1. Membrane integrity assays are particularly important as a measure of cellular damage,<br />

since there is evidence that some cationic nanoparticles, such as amine terminated<br />

dendrimers, exhibit toxic effects <strong>by</strong> disrupting the cell membrane. 87 Examples of<br />

assays that measure membrane integrity include the trypan blue exclusion assay and<br />

lactate dehydrogenase (LDH) leakage assay, which measures the presence of LDH<br />

released into the media through cell lysis. 88,89 The LDH leakage assay was selected<br />

because of its sensitivity and suitability for the high-throughput, 96-well plate format.<br />

2. Examples of assays which measure metabolic activity include tetrazolium dye reduction,<br />

ATP, and 3H-thymidine incorporation assays. The 3-(4,5-dimethyl-2-thiazolyl)-2,5diphenyl-2H-tetrazolium<br />

bromide (MTT) reduction assay was chosen for measurement<br />

of metabolic activity in the assay cascade, since it does not use radioactivity, and<br />

historically has been proven sensitive and reliable. MTT is a yellow water-soluble<br />

tetrazolium dye that is metabolized <strong>by</strong> live cells to water insoluble, purple formazan<br />

crystals. The formazan can be dissolved in DMSO and quantified <strong>by</strong> measuring the<br />

absorbance of the solution at 550 nm. Comparisons between the spectra of samples<br />

from nanoparticle treated and untreated cells can provide a relative estimate of cytotoxicity.<br />

90<br />

The MTT assay requires a solubilization step that is not required for the newer<br />

generation of tetrazolium dyes that form water-soluble formazans (e.g., XTT).<br />

However, these analogs require an intermediate electron acceptor that is often unstable,<br />

adding to assay variability. Furthermore, the net negative charge of these newer analogs<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


118<br />

limits cellular uptake, resulting in extracellular reduction. 91 MTT, with a net positive<br />

charge, readily crosses cell membranes and is reduced intracellularly, primarily in the<br />

mitochondria. Because nanoparticles have been shown to interact with cell membranes<br />

and could potentially interfere with the reduction of the newer generation analog via<br />

trans-plasma membrane electron transport, the traditional MTT assay would appear to be<br />

a better choice to assess cellular viability in nanoparticle cytotoxicity experiments.<br />

Analytes that are antioxidants, or are substrate/inhibitors of drug efflux pumps, have<br />

been shown to interfere with the MTT assay. 92,93 Functionalized fullerenes, which have<br />

not identified as efflux pump inhibitors or substrates, but do possess potent antioxidant<br />

activity, have been observed in our laboratory to cause MTT assay interference, resulting<br />

in enhanced MTT reduction and overestimation of cell viability (unpublished data).<br />

3. Loss of monolayer adherence to plating surfaces is often used as a marker of cytotoxicity.<br />

Monolayer adherence is commonly measured <strong>by</strong> staining for total protein, following<br />

fixation of adherent cells. This simple assay is often a very sensitive indicator of loss of<br />

cell viability. 55 The sulforhodamine B total protein staining assay was selected for<br />

determination of monolayer adherence. Advantages of this assay include the ability to<br />

store the fixed, stained microplates for extended periods prior to measurement, making<br />

the assay especially suitable for high throughput. 94<br />

4. Cell cycle analysis is conducted using propidium iodide staining of DNA and flow<br />

cytometry. 95 Flow cytometric can be used as a screening test for toxicity of chemicals.<br />

This method can determine the effect of nanoparticle treatment on cell cycle progression,<br />

as well as cell death. Cell cycle effects have been shown for a variety of nanoparticles.<br />

For instance, carbon nanotubes have been shown to cause G1 cell cycle arrest in human<br />

embryonic kidney cells, with a corresponding decrease in expression of G1-associated<br />

cdks and cyclins. 96<br />

7.3.4 OXIDATIVE STRESS<br />

Nanotechnology for Cancer Therapy<br />

The generation of free radicals <strong>by</strong> nanomaterials is well documented. 97,98 In most cases, the studied<br />

material was of ambient or industrial origin (quartz, carbon black, metal fumes, and diesel exhaust<br />

particles). However, engineered nanomaterials, such as fullerenes and polystyrene nanoparticles,<br />

have been shown to generate oxidative stress as well. 40,99,100 Lovric et al., for example, determined<br />

ROS to play an important role in cytotoxicity of quantum dots that have lost their protective<br />

coating. 101 The unique surface chemistries, large surface area, and redox active or catalytic<br />

contaminants (e.g., metals, quinones) of nanoparticles can facilitate ROS generation. 102 For<br />

example, fullerenes can perform electron transfer (phase-I pathway) or energy transfer (phase-II<br />

pathway) reactions with molecular oxygen following photoexcitation, 44 resulting in the formation<br />

of the superoxide anion radical or singlet oxygen, respectively. The superoxide anion radical can<br />

then undergo further reactions, such as dismutation and Fenton chemistry, to generate additional<br />

ROS species (e.g., % OH), resulting in cellular injury (see Scheme 7.1). 103 Evidence of fullereneinduced<br />

oxidative stress includes lipid peroxidation in the brains of exposed fish and treated rat liver<br />

microsomes. 40,104 Additional biomarkers of oxidative stress include a decrease in the reduced<br />

glutathione/oxidized glutathione ratio (GSH/GSSG), DNA fragmentation, and protein<br />

carbonyls. 105<br />

Biomarkers of nanoparticle-induced oxidative stress measured in our laboratory include ROS,<br />

lipid peroxidation products, and GSH/GSSG ratio. The fluorescent dichlorodihydroflourescein<br />

(DCFH) assay is used for measurement of ROS, such as hydrogen peroxide. 106 DCFH-DA is a<br />

ROS probe that undergoes intracellular deacetylation, followed <strong>by</strong> ROS-mediated oxidation to a<br />

fluorescent species, with excitation 485 nm and emission 530 nm. DCFH-DA can be used to<br />

measure ROS generation in the cytoplasm and cellular organelles, such as the mitochondria. The<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 119<br />

2GSH<br />

NADP +<br />

(1) ROS Generation<br />

hv<br />

GR<br />

2H<br />

GSSG +H2O GSHPx<br />

SOD<br />

Fullerene<br />

+<br />

–<br />

O2 O2 + e<br />

–<br />

O2 –<br />

NADPH + H +<br />

(2) Loss of GSH Homeostasis<br />

Fe ++<br />

GSH<br />

O 2 –<br />

O 2 + OH – + HO<br />

thiobarbituric acid reactive substances (TBARS) assay is used for measurement of lipid peroxidation<br />

products, such as lipid hydroperoxides and aldehydes. A molondialdehyde (MDA)<br />

standard curve is used for quantitation. MDA, a lipid peroxidation product, combines with thiobarbituric<br />

acid in a 1:2 ratio to form a fluorescent adduct, that is measured at 521 nm (excitation)<br />

and 552 nm (emission). TBARS are expressed as MDA equivalents. 107 The dithionitrobenzene<br />

(DTNB) assay is used for evaluation of glutathione homeostasis. In the DTNB assay, reduced GSH<br />

interacts with 5,5 0 -dithiobis(2-nitrobenzoic acid) (DTNB) to form the colored product 2-nitro-5thiobenzoic<br />

acid, which is measured at 415 nm, and GSSG. GSSG is then reduced <strong>by</strong> glutathione<br />

reductase to form reduced GSH, which is again measured <strong>by</strong> the preceding method. Pretreatment<br />

with thiol-masking reagent, 1-methyl-4-vinyl-pyridinium trifluoromethane sulfonate, prevents<br />

GSH measurement, resulting in measurement of GSSG alone. 108<br />

7.3.5 APOPTOSIS AND MITOCHONDRIAL DYSFUNCTION<br />

Nanoparticle-induced cell death can occur <strong>by</strong> either necrosis or apoptosis, processes that can be<br />

distinguished both morphologically and biochemically. Morphologically, apoptosis is characterized<br />

<strong>by</strong> perinuclear partitioning of condensed chromatin and budding of the cell membrane to<br />

form apoptotic bodies, whereas necrosis is characterized <strong>by</strong> cellular swelling (oncosis) and blebbing<br />

of the cell membrane. 109 In vitro studies have demonstrated the ability of nanoparticles, such<br />

as dendrimers and carbon nanotubes, to induce apoptosis. 110–112 In vitro exposure of macrophagelike<br />

mouse RAW 264.7 cells to cationic dendrimers led to apoptosis confirmed <strong>by</strong> morphological<br />

observation and the evidence of DNA cleavage. 112 Pretreatment of cells with a general caspase<br />

inhibitor (zVAD-fmk) reduced the apoptotic effect of the cationic dendrimer. 112 Apoptosis has also<br />

been observed in cultured human embryonic kidney cells (HEK293) and T lymphocytes treated<br />

CAT<br />

H 2 O 2 + O 2<br />

H 2 O 2<br />

2H 2O + O 2<br />

(3) Cellular Damage Lipid Peroxidation Protein/DNA Damage<br />

SCHEME 7.1 (1) Photoexcited fullerenes can perform electron-transfer reactions with molecular dioxygen to<br />

form the superoxide anion radical (O $K<br />

2 ). Superoxide can then undergo superoxide dismutase (SOD)-catalyzed<br />

dismutation to hydrogen peroxide (H 2O 2). H 2O 2 is a substrate for catalase (CAT)-and glutathione peroxidase<br />

(GSHPx)-catalyzed detoxification reactions. (2) The oxidation of glutathione (GSH) to form oxidized glutathione<br />

(GSSG) during detoxification of H2O2 can result in a loss of glutathione homeostasis. GSH can be<br />

regenerated <strong>by</strong> glutathione reductase (GR). (3) Alternatively, hydrogen peroxide can undergo transition metal<br />

(Fe CC )-catalyzed Fenton chemistry to form the highly reactive hydroxyl radical (HO % ) that is capable of<br />

initiating lipid peroxidation and DNA/protein oxidation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


120<br />

Nanotechnology for Cancer Therapy<br />

with single walled carbon nanotubes, and in MCF-7 breast cancer cells treated with quantum<br />

dots. 101,110,113<br />

Apoptosis in mammalian cells can be initiated <strong>by</strong> four potential pathways: (1) mitochondrial<br />

pathway, (2) Death receptor-mediated pathway, (3) ER-mediated pathway, and (4) Granzyme<br />

B-mediated pathway. 114 Our laboratory has focused on caspase-3 activation in liver and kidney<br />

cells as a biomarker of apoptosis, since this a downstream event in all the classical apoptotic<br />

signaling pathways and can be measured using a fluorometric protease assay. This assay quantifies<br />

caspase-3 activation in vitro <strong>by</strong> measuring the cleavage of DEVD-7-amino-4-trifluoromethyl<br />

coumarin (AFC) to free AFC that emits a yellow-green fluorescence (l maxZ505 nm). 115 This<br />

initial apoptosis screen can then be followed <strong>by</strong> additional analysis, as cellular morphology<br />

studies using nuclear staining techniques to detect perinuclear chromatin, or agarose gel electrophoresis<br />

to detect DNA laddering. 116<br />

Evidence supports a role for ROS in generation of the mitochondrial permeability transition via<br />

oxidation of thiol components of the permeability transition pore complex. 117 As discussed in the<br />

preceding sections, nanoparticles have been shown to induce oxidative stress, and thus this ROSmediated<br />

pathway for induction of the mitochondrial permeability transition is a plausible apoptotic<br />

mechanism for nanomaterials. For instance, ambient ultrafine particulates have been shown to<br />

translocate to the mitochondria of RAW 264.7 murine macrophage cells, cause structural<br />

damage, and altered mitochondrial permeability. 98 A subsequent study demonstrated that mitochondrial<br />

dysfunction and apoptosis in the RAW 264.7 cells could be induced <strong>by</strong> polar compounds<br />

fractionated from ultrafine particles, suggesting that the mitochondrial dysfunction caused <strong>by</strong><br />

ultrafines was the result of redox cycling of quinone contaminants on the surface of the particle. 118<br />

This link between oxidative stress, mitochondrial dysfunction, and apoptosis has also been<br />

observed for man-made nanoparticles. For example, metal and quantum dot engineered nanoparticles<br />

have both been shown to induce oxidative stress, mitochondrial dysfunction, and apoptosis in<br />

various in vitro models. 101,119 Water-soluble, derivatized fullerenes, which have been shown to<br />

accumulate in the mitochondria of HS 68 human fibroblast cells, have also been shown to induce<br />

apoptosis in U251 human glioma cells. 120,121 While this derivatized fullerene-induced apoptosis in<br />

the glioma cell line did not involve oxidative stress, mitochondrial dysfunction was not measured<br />

and cannot be ruled out. Mitochondrial dysfunction and apoptosis have also been observed in a<br />

human gastric carcinoma cell line exposed to chitosan nanoparticles. 122 Taken together, these<br />

observations support a role for mitochondrial dysfunction and oxidative stress in nanoparticleinduced<br />

apoptosis. Apart from apoptosis, mitochondrial dysfunction has long been associated<br />

with necrotic cell death, and represents a potential necrotic mechanism of nanoparticle-induced<br />

injury as well. 123<br />

Mitochondrial dysfunction can result from several mechanisms in addition to opening of the<br />

permeability transition pore complex, including uncoupling of oxidative phosphorylation, damage<br />

to mitochondrial DNA, disruption of the electron transport chain, and inhibition of fatty acid<br />

b-oxidation. 124 Methods used to detect mitochondrial dysfunction include measurement of<br />

ATPase activity (via luciferin–luciferase reaction), oxygen consumption (via polarographic technique),<br />

morphology (via electron microscopy), and membrane potential (via fluorescent probe<br />

analysis). 125 Our laboratory measured loss of mitochondrial membrane potential in rat hepatic<br />

primaries, and Hep-G2 and <strong>LLC</strong>-PK1 cell lines, using the 5,5 0 ,6,6 0 -tetrachloro-1,1 0 ,3,3 0 -tetraethylbenzimidazolcarbocyanine<br />

iodide (JC-1) assay, which is a convenient assay that does not require<br />

mitochondrial isolation or use specialized equipment. 126 This fluorescent dye partitions into the<br />

mitochondrial matrix as a result of the membrane potential. Concentration of JC-1 in the matrix<br />

results in aggregation that fluoresces at 590 nm (red). Upon loss of membrane potential, the dye<br />

dissipates from the matrix and can be measured, in its monomer state at emission 527 nm (green).<br />

The proportion of green to red fluorescence reflects the degree of mitochondrial membrane<br />

depolarization.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 121<br />

7.3.6 PROTEOMICS AND TOXICOGENOMICS<br />

Proteomics and toxicogenomics are useful tools for identifying the mechanisms underlying<br />

toxicity. 127 Using gel electrophoresis in combination with MS identification, or gene microarray<br />

technology, the expression of specific pathway-responsive genes, such as Phase-II enzymes for<br />

oxidative stress, or cytokines for inflammation, can be identified. The delineation of these toxic<br />

pathways could help further refine the in vitro and in vivo study of nanomaterials, potentially<br />

leading to the development of novel biomarkers that could then be used in clinical and occupational<br />

toxicology studies. Proteomic and genomic research on biomedically relevant nanomaterials is<br />

presently underway, using a series of human hepatocyte, kidney, and immunological primary cells.<br />

7.4 IN VIVO PHARMACOKINETIC AND TOXICOLOGICAL ASSESSMENTS<br />

As is the case with any NCE, a thorough understanding of the properties that govern biocompatability<br />

is necessary to allow transition of nanomaterials to human clinical trials. Although in vitro<br />

toxicology studies can be informative, the obvious caveat is that phenomenon observed in vitro may<br />

not materialize in vivo due to differences in biological response or nanoparticle concentrations.<br />

Therefore, nanoparticle safety and therapeutic efficacy can only be definitively assessed <strong>by</strong> rigorous<br />

in vivo testing. This phase is guided in part <strong>by</strong> insights obtained from the physicochemical and<br />

in vitro characterization programs.<br />

The primary goal of in vivo studies is to evaluate nanomaterials’ pharmacokinetics, safety, and<br />

efficacy (see Table 7.1) in the most appropriate animal models. Preclinical toxicological and<br />

pharmacokinetic studies are conducted in accordance with the FDA regulatory guidance for IND<br />

and IDE submission. It is not within the scope of this chapter to review this regulatory guidance;<br />

instead, the reader is directed to the regulatory chapter of this text. While it is generally agreed that<br />

the current in vivo pharmacological and toxicological endpoints used for devices and small<br />

molecule drugs should be appropriate in assessing the safety and efficacy of biomedical nanomaterials,<br />

the qualities of nanomaterials that lend themselves to biomedical application, such as<br />

TABLE 7.1<br />

In Vivo Pharmacological and Toxicological Assessment of Nanomatersials<br />

Category Assessment<br />

Initial disposition study Tissue distribution<br />

Clearance mechanisms<br />

Half-life<br />

Systemic exposure (plasma AUC)<br />

Immunotoxicity 28-day screen<br />

Immunogenicity (repeat-dose toxicity study)<br />

Hypersensitivity<br />

Immunostimulation<br />

Immunosuppression<br />

Dose-range finding toxicity NOAEL<br />

STD10<br />

Good laboratory practices studies PK/ADME<br />

Expanded single dose acute toxicity<br />

Repeated dose toxicity<br />

Efficacy Targeting<br />

Therapeutic<br />

Imaging<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


122<br />

Nanotechnology for Cancer Therapy<br />

macromolecular structure and polydispersity, could also be problematic with respect to preclinical<br />

characterization, as will be discussed below. Early efforts are focused on identifying and standardizing<br />

the analytical and toxicological methodologies that are unique to nanoparticle<br />

preclinical characterization.<br />

Several factors can influence the clinical viability of a new cancer diagnostic or therapeutic<br />

agent, aside from economic feasibility. These include: (1) demonstrated advantage over the current<br />

market standard; (2) appropriate administration route/schedule/elimination half-life; and (3)<br />

favorable safety profile.<br />

1. The in vivo characterization phase includes an assessment of nanoparticle imaging<br />

and/or therapeutic efficacy in animal models that most closely approximate the human<br />

disease state. For instance, targeting efficacy is addressed <strong>by</strong> comparing a nanoparticle<br />

distribution profile with that of a nontargeted nanoparticle from the same class; however,<br />

this approach may be prone to ambiguities due to passive targeting via the EPR effects.<br />

For those particles with imaging components, the signal enhancement and tissue distribution<br />

profile is monitored using the appropriate magnetic resonance, ultrasound, optical,<br />

or positron emission tomography imaging instrumentation. In all cases, the approach will<br />

be compared against the present market standard to provide comparable data regarding<br />

efficacy, pharmacokinetics, and safety.<br />

2. Animal studies of nanoparticles have rarely utilized the oral administration route, but<br />

those that have used that route have demonstrated poor intestinal absorption. For<br />

example, 98% of orally administered, PEG-functionalized fullerenes are eliminated in<br />

the feces within 48 h. Oral bioavailability of polystyrene nanoparticles were similarly<br />

poor, with less than 7% of 3000-, 1000-, 100-, 50-, and 3-nm sized particles absorbed. 128<br />

Due to this extremely low nanoparticle oral bioavailability, the majority of biomedical<br />

nanoparticle formulations encountered undoubtedly will be intended for parenteral<br />

administration. Since diagnostic and chemotherapeutic regimens are typically of short<br />

duration, the inconvenience of intravenous administration does not appear to be a significant<br />

hurdle for eventual clinical transition.<br />

To be a successful diagnostic or therapeutic agent, nanoparticles should be eliminated<br />

from the body in a reasonable timeframe. However, studies have shown that optimum<br />

passive targeting of tumors, <strong>by</strong> the EPR effect, also requires that nanoparticle agents<br />

remain in the systemic circulation for prolonged periods. 129 Therefore, there must be a<br />

balance between systemic residence and clearance. At present, avoidance of the RES<br />

system and eventual urinary clearance appears to be a formidable obstacle for many of<br />

the current approaches, such as iron oxide MRI contrast agents and lipidic nanoparticle<br />

drug delivery agents, which have been shown to undergo capture <strong>by</strong> organs of the RES<br />

and remain for extended periods. 130,131 The primary mechanism of nanoparticle clearance,<br />

as discussed previously, appears to be glomerular filtration, which is governed <strong>by</strong><br />

charge, molecular weight, and degree of protein binding. 132,133 A good case study of the<br />

importance of molecular weight in mediating glomerular filtration, <strong>by</strong> Gillies and<br />

colleagues, demonstrated that dendrimer–polyethylene oxide complexes greater than<br />

40 kDa were cleared less readily than lower molecular weight species. Because timely<br />

clearance is an important drug attribute, accurate and thorough disposition studies<br />

are required.<br />

The single greatest obstacle for nanoparticle disposition studies is analytical<br />

methodology to quantitate nanoparticle concentrations in biological matrices, such as<br />

plasma and tissues. Due to their macromolecular and polydispersed nature, nanomaterials<br />

do not lend themselves to quantitation <strong>by</strong> traditional methods, such as HPLC and<br />

LC/MS, and may require alternative methods such as radiolabeling and scintillation<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 123<br />

counting. Since many biomedical nanoparticles will be multifunctional, and have<br />

imaging components, the imaging functionality of the particle could potentially be<br />

used for in vivo quantitation. In cases where no imaging component is present, the<br />

nanoparticles could be tagged with an appropriate probe to allow for imaging quantitation.<br />

The labeling method utilized may alter the surface properties of a<br />

nanoparticle, and thus affect the tissue distribution profile; comparison of alternative<br />

labeling methods would help identify consensus behavior. In any event, the use of<br />

imaging would first require validation using traditional methods such as LC/MS or<br />

scintillation to ensure image intensity could correlate to nanoparticle concentration in<br />

a linear fashion. Since many nanomaterials are electrondense, electron microscopy might<br />

also be used for tracking tissue distribution.<br />

3. The objectives of the preclinical toxicological studies are to identify target organs of<br />

toxicity and to aid in the selection of starting doses for phase-I human clinical trials.<br />

Toward this end, toxicity studies seek to determine dose ranges causing (1) no adverse<br />

effects (NOAEL) and (2) life-threatening toxicity (i.e., severe toxic dose 10% [STD 10]).<br />

Studies are performed in two mammalian species, rodent and nonrodent, with rats the<br />

preferred rodent species, since they exhibit the greatest concordance with human toxicities.<br />

134 Nanoparticle formulations are administered according to the intended clinical<br />

treatment cycle, with regard to schedule, duration, route, and formulation. Necropsy,<br />

performed on animals showing signs of morbity during the study and at study termination,<br />

includes comprehensive hematology, histopathology, and clinical chemistry (see<br />

Figure 7.3). Several preclinical studies suggest a key role for reticuloendothelial organs,<br />

such as liver, kidney, and bone marrow, in the uptake of nanoparticles from the systemic<br />

circulation. 67,69 Therefore, these organs should receive special attention with regard to<br />

functional and histopathological evaluation. A review of the limited in vivo safety data<br />

available for nanoparticles supports the scrutinizing of these tissues, as there are several<br />

examples of RES organ injury, including hepatotoxicity and nephrotoxicity. For<br />

example, intravenous administration of cationic PAMAM dendrimers at low doses has<br />

Brain<br />

Histopathology<br />

Pancreas<br />

Lymph node Esophagus<br />

Thyroid<br />

Trachea<br />

Pituitary<br />

Heart<br />

Thymus<br />

Gall bladder<br />

Spleen<br />

Lung<br />

Ileum<br />

Rectum<br />

Cecum<br />

Colon<br />

Lymph node Epididymis<br />

Prostate Seminal vesicle<br />

Urinary bladder Uterus<br />

Hardian gland Nasal sections<br />

Femur<br />

Vertebra<br />

Mammary gland Skin/subcuits<br />

BUN<br />

GGT<br />

total protein<br />

A/G<br />

Chloride<br />

Clinical Chemistry<br />

AST<br />

GLUC<br />

Albumin<br />

Sodium<br />

Salivary gland<br />

Parathyroid<br />

Adrenal<br />

Kidney<br />

Liver<br />

Duodenum<br />

Stomach<br />

Jejunum<br />

Ovary<br />

Testis<br />

Eye<br />

Femur<br />

Spinal cord<br />

Tongue<br />

ALT<br />

Creatinine<br />

Globulin<br />

Potassium<br />

Hematology<br />

Erthrocyte count (RBC)<br />

Hemoglobin (HGB)<br />

Hematocrit (HCT)<br />

Mean corpuscular volume (MCV)<br />

Mean corpuscular hemoglobin (MCH)<br />

Mean corpuscular hemoglobin conc. (MCHC)<br />

Platelet count (Plate)<br />

Reticulocyte count (RETIC)<br />

Total leukocyte count (WBC)<br />

Differential leukocyte count<br />

Nucleated red blood cell count<br />

FIGURE 7.3 Panel of tissues and chemistries to be assessed for comprehensive toxicology.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


124<br />

been shown to cause liver injury, as determined <strong>by</strong> histopathology and elevations in<br />

serum alanin aminotransferase when administered intravenously to mice; larger doses of<br />

the same cationic dendrimer were lethal to 100% of the mice. 135 Nephrotoxicity and<br />

hepatotoxicity, as determined <strong>by</strong> histopathology and serum enzyme markers, have both<br />

been observed in mice treated orally with nano-zinc. 72,84<br />

One of the potential advantages of nanoparticle drug formulations is an improved safety profile<br />

as a result of targeted therapy or elimination of toxic solubilization agents. For example, Baker and<br />

colleagues have shown that methotrexate-conjugated PAMAM dendrimers containing a folate<br />

receptor targeting ligand are more efficacious, and less toxic, than unformulated methotrexate<br />

against a murine human epithelial cancer model. 34 Abraxane is an example of a nanoparticle<br />

formulation of paclitaxel, presently on the market, that takes advantage of the solubilizing effect<br />

of albumin, eliminating the need for the toxic vehicle Cremophor EL. 136 Phase-III studies have<br />

shown enhanced therapeutic response of abraxane compared to Cremophor-formulated paclitaxel,<br />

while side effects, such as myelosuppression and peripheral neuropathy, were significantly reduced.<br />

As discussed above, nanoparticle drug formulations are compared against the unformulated drug to<br />

determine if the improvement in safety profile is realized.<br />

7.5 IMMUNOTOXICITY<br />

Nanotechnology for Cancer Therapy<br />

A growing body of evidence suggests that immunotoxicity provides a considerable contribution to<br />

onset and development of various disorders, including cancer and autoimmune diseases. 137–139<br />

Nevertheless, it was not until recently that this relatively new field of toxicology emerged as an<br />

important interface between the fields of novel drug design and pharmacology. Recognition of<br />

immunosuppressive properties of new pharmaceuticals during early drug development phase is<br />

very important to eliminate potentially dangerous substances from the drug pipelines. For<br />

example, treatment of patients diagnosed with Crohn’s disease and rheumatoid arthritis with infliximab<br />

and etanercept (both drugs represent neutralizing anti-TNF antibodies) resulted in increased<br />

incidence of tuberculosis and histoplasmosis. 140–143 Although these data did not result in withdrawal<br />

of any of the products from the pharmaceutical market, they helped initiate the strategy of preparing<br />

patients for anti-TNF therapy <strong>by</strong> screening for, and treatment of, latent tuberculosis prior to administration<br />

of anti-TNF medications. Immunosuppression caused <strong>by</strong> pharmaceuticals can also lead to<br />

the development of lymphomas and acute leukemia. 144–146 Undesirable immunostimulation caused<br />

<strong>by</strong> pharmacological intervention include immunogenicity, hypersensitivity, and increased risk of<br />

autoimmune response. The standard toxicology endpoints employed for safety assessment of new<br />

pharmaceuticals primarily rely on clinical chemistry and histopathological evaluation of immune<br />

organs and were developed several decades ago. Currently, there is an increasing demand for<br />

the development of new methods for immunotoxicity assessment because of drug candidates’<br />

more complex structure as well as the application of new technologies in their manufacturing.<br />

The introduction of new molecular and immune cell biology methods into the immunotoxicology<br />

assessment framework is not a trivial and straightforward process. It requires not only scrupulous<br />

validation and standardization of the new techniques but also demonstration of the physiological<br />

relevance for the proposed battery of assays. These processes are expensive, time-consuming, and<br />

necessitate cooperation across the various pharmaceutical industry players. Unlike traditional drugs,<br />

multifunctional nanomaterials combine both chemistry-based and biotechnology-derived components,<br />

and therefore their characterization using standard methodologies requires adjustments<br />

and/or modification of classical experimental protocols. Below we will attempt to summarize data<br />

on critical aspects of immunotoxicological evaluation of nanomaterials and examine challenges in<br />

the application of standard methodologies for the assessments of nanoparticle safety to the<br />

immune system.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 125<br />

7.5.1 APPLICABILITY OF STANDARD IMMUNOLOGICAL METHODS FOR NANOPARTICLE EVALUATION<br />

AND CHALLENGES SPECIFIC TO NANOPARTICLE CHARACTERIZATION<br />

Immunotoxicological evaluation of new drug candidates includes studies on both immunosuppression<br />

and immunostimulation, and is applicable to nanomaterials intended for use as drug candidates<br />

and/or drug delivery platforms. Short-term in vitro assays are being developed to allow for quick<br />

evaluation of nanoparticles’ biocompatibility. The in vitro immunotoxicity assay cascade includes<br />

the following methods: analysis of plasma protein binding <strong>by</strong> two-dimensianol polyacrylamide gel<br />

electrophoresis (PAGE), hemolysis, platelet aggregation, coagulation, complement activation,<br />

colony forming unit-granulocyte macrophage (CFU-GM), leukocyte proliferation, phagocytosis,<br />

cytokine secretion <strong>by</strong> macrophages, chemotaxis, oxidative burst, and evaluation of cytotoxic<br />

activity of NK cells. 147 In addition to these methods, our in vitro tests include sterility assessment<br />

based on pyrogen contamination test (L-amebocyte lysate (LAL) assay) and evaluation of microbiological<br />

contamination. The assay cascade is based on several regulatory documents<br />

recommended buy the FDA for immunotoxicological evaluation of new investigational drugs,<br />

medical devices, and biotechnology, derived pharmaceuticals, 148–152 as well as on ASTM and<br />

ISO standards developed for characterization of blood contact properties of medical<br />

devices. 153–155 The aim of the in vitro immunoassay cascade is to provide quick evaluation of<br />

nanomaterials of interest prior to initiation of more thorough in vivo studies. Challenges specific for<br />

immunotoxicity assessment of nanoparticulate materials are summarized below.<br />

7.5.1.1 Blood Contact Properties<br />

One important aspect of nanoparticle used for medical applications is the assurance that they will<br />

not cause toxicity to blood elements when injected into a patient.<br />

Hemolysis (i.e., damage to red blood cells) can lead to life-threatening conditions such as<br />

anemia, hypertension, arrhythmia, and renal failure. In our laboratory we have developed a protocol<br />

to evaluate hemolytic properties of nanoparticles based on the existing ASTM International standard<br />

used to characterize other materials. 147,156 We have identified several problems when applying<br />

existing protocol for nanoparticles characterization. For example, colloidal gold nanoparticles with<br />

size 5–50 nm have absorbance at 535 nm, which overlaps with the assay wavelength of 540 nm.<br />

Removal of these particles <strong>by</strong> centrifugation was required prior to sample evaluation for the<br />

presence of plasma-free hemoglobin to avoid false-positive results. Though it worked well for<br />

gold particles with size 10–50 nm, centrifugation may be problematic for other nanoparticles.<br />

For example, small colloidal gold particles with a size of 5 nm require higher centrifugation<br />

force to be removed from the supernatant. Hemoglobin has a size of 5 nm; therefore, one cannot<br />

exclude the possibility that ultracentrifugation of supernatant may pellet hemoglobin along with the<br />

gold particles and thus result in a false-negative result. Ultracentrifugation is not feasible for<br />

fullerenes or dendrimer particles. Analysis of polystyrene particles of 20, 50, and 80 nm revealed<br />

another complication. We found that particle preparation damages red blood cells; the damage is<br />

caused <strong>by</strong> the surfactant used during particle manufacturing, and is detected for 20- and 50-nm<br />

particles. For 80-nm particles, the hemolysis assay showed false-negative results due to the adsorption<br />

of hemoglobin <strong>by</strong> the particles. When we applied dialysis to remove surfactant, 50-nm particles<br />

revealed same phenomenon as 80-nm particles, i.e., they caused hemolysis, but adsorbed hemoglobin<br />

resulting in a false-negative result. Another potential problem with the application of<br />

standard hemolysis protocol for nanoparticles characterization is that metal-containing particles<br />

may oxidize hemoglobin and result in a change in assay OD responses. Therefore, assay procedures<br />

may require slight modifications depending on the particle type. In general, inclusion of particle test<br />

samples without blood allowed for quick assessment of the potential particle interference with the<br />

assay. In some instances, deduction of the result generated for blood-free particle control from that<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


126<br />

Nanotechnology for Cancer Therapy<br />

obtained for blood-plus particle sample was possible and allowed estimation of particle potential to<br />

cause damage to erythrocytes.<br />

Blood coagulation. Blood coagulation may be affected <strong>by</strong> nanomaterials. For example, modification<br />

of surface chemistry has been shown to improve immunological compatibility at the<br />

particle–blood interface: application of polyvinyl chloride resin particles resulted in 19G4%<br />

decrease in platelet count, indicating platelet adhesion/aggregation and increased blood coagulation<br />

time; the same particle preparation coated with PEG affected neither the platelet count nor elements<br />

of coagulation cascade. 157 Similarly, folate-coated and PEG-coated Gd nanoparticles did not<br />

aggregate platelets or activate neutrophils. 158 The evaluation of nanoparticle effects on blood<br />

coagulation includes studies on platelet function and coagulation factors. The in vitro cascade<br />

includes platelet aggregation assay and four coagulation assays measuring prothrombin time,<br />

activated partial thromboplastin time, thrombin time, and reptilase time.<br />

Interaction with plasma proteins. High-resolution, 2D PAGE may be the method of choice to<br />

investigate plasma protein adsorption <strong>by</strong> nanoparticles. 2D PAGE has been used in several labs to<br />

isolate and identify plasma proteins adsorbed on the surface of stealth polycyanoacrylate particles 8 ,<br />

liposomes 159,160 , solid lipid 161 , and iron oxide nanoparticles. 162 Proteins commonly identified on<br />

several types of nanomaterials include antithrombin, C3 component of complement, alpha-2macroglobulin,<br />

haptoglobin, plasminogen, immunoglobulins, albumin, fibrinogen, apolipoprotein,<br />

and transthyretin; albumin, immunoglobulins, and fibrinogen are the most abundant. Studies using<br />

this approach revealed that surface chemistry is important for protein adsorption. For example,<br />

Peracchia et al. demonstrated that coating with PEG results in approximately a fourfold reduction in<br />

protein binding <strong>by</strong> polycyanoacrylate particles. 163 Gessner et al. prepared polystyrene latex model<br />

nanoparticles with different surface charges. This study demonstrated that increasing surface charge<br />

density results in a quantitative increase in plasma protein adsorption, but did not show significant<br />

differences in the qualitative composition of the absorbed protein mixture. 164<br />

One of the most important step in this procedure is the separation of particles from plasma after<br />

incubation is complete. Ultracentrifugation was shown to be successful for isolation of iron oxide,<br />

solid lipid particles, and some polymer-based particles. 8,160,162 Gel filtration was applicable to<br />

liposomes, 159 solid lipid, and iron oxide nanoparticles. Thode and colleagues compared four<br />

methods for isolation of iron oxide particles, i.e., ultracentrifugation, static filtration, magnetic<br />

separation, and gel filtration. Depending on the method used for particle separation from bulk<br />

plasma, different quantities of the same proteins and different species of proteins were identified<br />

on the particles of the same size and surface chemistry. 162 For example, albumin was the predominant<br />

protein if static filtration and gel filtration were employed, while small quantities of this<br />

protein were found after ultracentrifugation; it was almost undetectable when magnetic separation<br />

was used. There was no difference in isolation of fibrinogen among the four methods. Comparable<br />

quantities of IgG gamma-chain were isolated using ultracentrifugation, static filtration, and<br />

magnetic separation, while gel filtration appeared to be inefficient in isolation of this protein. 162<br />

Attention has to be paid to the sample preparation to avoid artificial protein adsorption due to<br />

desorption during the separation, for example. Other critical steps are the number of washes to<br />

remove an excess of bulk plasma, the type of wash buffer, and a buffer to dislodge protein from the<br />

particle surface. In our lab we also found that using polypropylene low-retention tubes and pipette<br />

tips is crucial for isolation of particle-specific proteins (unpublished data).<br />

Complement activation and phagocytosis. Following intravenous administration, nanoscale<br />

drug carriers may suffer a drawback in that they may be taken up <strong>by</strong> cells of the mononuclear<br />

phagocytic system. Consequently, such uptake facilitates clearance of nanoparticulate carriers and<br />

associated drugs, thus leading to a decrease in drug efficacy. 165 The initial adsorption of plasma<br />

proteins such as components of complement and immunoglobulins promotes nanoparticles clearance.<br />

166,167 Therefore, the investigation of a nanoparticle’s ability to interact with and activate a<br />

complement and uptake <strong>by</strong> mononuclear cells seems to be one of the key assays in the preclinical<br />

characterization cascade. Classical immunoassays used to evaluate complement activation, such as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 127<br />

the total hemolytic complement assay (CH50) and the alternative pathway (rabbit CH50 or<br />

APCH50), are based on the hemolysis of rabbit erythrocytes. These hemolytic assays can be<br />

used to measure functional activity of specific components of either pathway. The main challenge<br />

in applying these assays for nanomaterial characterization is the ability of nanoparticles per se to<br />

lyse RBCs, thus generating false-positive results. To overcome this limitation, the approach for<br />

evaluation of complement activation includes two techniques. One is a qualitative yes or no rapid<br />

screen for the presence of C3 cleavage products using western blot. The second assay is a quantitative<br />

evaluation of samples found positive at an initial screen for the presence of C4a, C3a, and<br />

C5a components of complement using a flow cytometry-based multiplex array.<br />

The difficulties with application of standard phagocytosis assay are: (1) light microscopy used<br />

in traditional phagocytosis assay 168 is not applicable to nanoparticles due to their smaller size; (2)<br />

when light microscopy is substituted with TEM, visualization of particle may be complicated since<br />

their size is similar to that of cell organelles resulting in ambiguous interpretation of TEM data, thus<br />

TEM is limited to electron dense metal containing particles (see Figure 7.4); (3) labeling of<br />

nanoparticles with fluorescent tags 9,169 is a superior approach for visualizing internalized particles,<br />

but should be avoided in preclinical tests as chemical attachment of fluorophore may create a new<br />

molecular entity with properties widely divergent from those of the original particle; (4) application<br />

of luminol to detect phagocytosed particles 8 provides an exceptional technique which can overcome<br />

all limitations listed above; however, it is not free of applicability reservations as well (e.g.,<br />

nanoparticles may interfere with activation of luminol once it is internalized, etc.).<br />

CFU-GM assays. CFU-GM assays allow for the evaluation of potential nanoparticles<br />

interference with growth and differentiation of bone marrow stem cells into granulocyte and<br />

macrophages. This assay may provide valuable information for development of anti-cancer nanotechnology<br />

platforms. Myelosuppression is a very common dose-limiting toxicity associated with<br />

the use of oncology cytotoxic drugs. Incorporation of such drugs into nanoparticle carriers targeted<br />

to specific cancer cells may help to reduce toxicity to normal tissues, including bone marrow, and<br />

should be considered during initial characterization of nanocarriers.<br />

(a)<br />

2μm<br />

(b) 0.2μm<br />

FIGURE 7.4 Study of internalization of 30-nm colloidal gold nanoparticles <strong>by</strong> murine macrophage cell line<br />

RAW 264.7 <strong>by</strong> TEM. (a) Analysis of single cell. (b) Zoom-in analysis of the selected area in the same cell.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


128<br />

7.5.2 IMMUNOGENICITY<br />

This issue of immune stimulation <strong>by</strong> pharmaceuticals came into the forefront when biotechnologyderived<br />

products, especially recombinant proteins, moved toward clinical trials. Today it is evident<br />

that the immune system can effectively recognize biological therapeutics as foreign substances and<br />

build up a multi-level immune response against them. A number of factors result in the immune<br />

system responding to the administration of a pharmacological product, such as structure, formulation,<br />

folding architecture, but also degradation <strong>by</strong>products. 170 In addition, the route of<br />

administration and the dosage were shown to influence the staging and the amplitude of the<br />

immune system’s response. In general, immune responses to biological products could be classified<br />

as benign in the sense that they affect only the pharmacological efficacy of the administered<br />

compound. The greatest concern is the robust immune response to certain biotechnology-derived<br />

products that are fraught with serious clinical consequences that may even result in a fatal outcome<br />

due to specific recognition and elimination of the patient’s endogenous growth factors critical for<br />

survival. 171–173 For example, in the case of thrombopoietin, the immune response may result in the<br />

production of neutralizing antibodies, causing inhibition of the endogenous thrombopoietin with<br />

subsequent development of thrombocytopenia. 173 The patient’s immune response to recombinant<br />

erythropoietin product Eprex w has been reported to induce pure red cell aplasia. 171,172,174 In the latter<br />

example, cross-reactivity tests indicated that antibodies generated against Eprex w could also neutralize<br />

other forms of erythropoietin products such as Epogen w , NeoRecormon w , and Aranesp w and<br />

suggested that antibodies are directed against some specific conformation of the erythropoietin active<br />

site. 170 Although incidence of the acute pure red cell aplasia remains relatively rare, the long-term<br />

implications are of great concern, as over half of patients who developed the auto-antibody remained<br />

transfusion dependent. The potential of using multifunctional nanoparticles for medical applications<br />

raises a key question of whether nanoparticle materials <strong>by</strong> themselves can induce an anti-nanoparticle<br />

immune response, stimulate allergic reactions, or trigger synthesis of nanopreparation-specific<br />

IgE. One can expect that the generation of antibodies to nanoparticles will ultimately affect only<br />

efficacy of the particle-based product. Of greatest concern will be the immune response to particles<br />

functionalized with growth factors, receptors or other biological molecules, which would result in the<br />

formation of antibodies, neutralizing the effect of these biological molecules and leading to potential<br />

exclusion of both particle-linked and endogenous proteins, akin to similar effects observed with<br />

biotechnology-derived pharmaceuticals. There are a limited number of studies on immunogenicity of<br />

nanomaterials. A few of them have shown that nanoparticles may both induce nanoparticles specific<br />

immune response and act as adjuvants. 175–178 Although preclinical animal studies may not be<br />

predictive to human immune response, available data described above do suggest that the immune<br />

system can recognize and build an immune response against some nanoparticles. Therefore, evaluation<br />

of nanoparticle antigenicity is seen as an important step during preclinical development. Other<br />

immunogenicity characterization should include evaluation of a nanoparticles’ ability to act as an<br />

adjuvant and to induce allergic reactions.<br />

7.6 CONCLUSIONS<br />

Nanotechnology for Cancer Therapy<br />

The urgency to eliminate the pain and suffering associated with cancer is fueling research into novel<br />

therapies at a blistering pace. Through exquisitely targeted and multifunctional approaches, nanotechnology<br />

in particular holds great promise for enhancing cancer therapy. This promise, however,<br />

will never be realized if the safety of the nanomaterial is not demonstrated to allay public concern<br />

and if the regulatory structure is not in place to allow proper evaluation of the science. Without such<br />

a framework, the return on investment will be uncertain and progress will undoubtedly be stalled.<br />

The effort to develop a standardized set of protocols to characterize nanomaterials and their<br />

biological effects will provide a foundation for future regulation and will lead to a body of knowledge<br />

that will guide the design of safer nanotechnology products.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 129<br />

The biologic activity and toxicity of nanoscale particles are dependent on many parameters not<br />

typically examined for conventional small molecule therapeutics: size, shape, surface chemistry,<br />

stability of outer coating, agglomeration state, etc., and many conventional properties, such as<br />

stability or biodistribution, must be analyzed using a very different set of protocols and/or instrumentation.<br />

Emerging data from studies on nanoparticles engineered for medical use will build<br />

toward a consensus of instrumentation and experimental methodology needed to reproducibly<br />

determine the pharmacology and safety of these novel products.<br />

The greatest challenge may not be in the development of new screening technologies, but in the<br />

ability to promulgate an accepted set of characterization protocols throughout the Nano Bio<br />

industry. The Nanotechnology Characterization Laboratory at the National Cancer Institute is<br />

working together with the FDA, NIST, and other regulatory and standards-setting organizations<br />

to establish the standard assays and technologies needed for timely delivery of safe and effective<br />

nanotechnology products.<br />

REFERENCES<br />

1. Furumoto, K., Nagayama, S., Ogawara, K., Takakura, Y., Hashida, M., Higaki, K., and Kimura, T.,<br />

Hepatic uptake of negatively charged particles in rats: Possible involvement of serum proteins in<br />

recognition <strong>by</strong> scavenger receptor, Journal of Controlled Release, 97, 133–141, 2004.<br />

2. Oberdorster, G., Toxicology of ultrafine particles: In vivo studies, Philosophical Transactions of the<br />

Royal Society of London Series A—Mathematical Physical and Engineering Sciences, 358,<br />

2719–2739, 2000.<br />

3. Ogawara, K., Yoshida, M., Higaki, K., Kimura, T., Shiraishi, K., Nishikawa, M., Takakura, Y., and<br />

Hashida, M., Hepatic uptake of polystyrene microspheres in rats: Effect of particle size on intrahepatic<br />

distribution, Journal of Controlled Release, 59, 15–22, 1999.<br />

4. Ogawara, K., Yoshida, M., Kubo, J., Nishikawa, M., Takakura, Y., Hashida, M., Higaki, K., and<br />

Kimura, T., Mechanisms of hepatic disposition of polystyrene microspheres in rats: Effects of serum<br />

depend on the sizes of microspheres, Journal of Controlled Release, 61, 241–250, 1999.<br />

5. Fenart, L., Casanova, A., Dehouck, B., Duhem, C., Slupek, S., Cecchelli, R., and Betbeder, D.,<br />

Evaluation of effect of charge and lipid coating on ability of 60-nm nanoparticles to cross an in vitro<br />

model of the blood–brain barrier, Journal of Pharmacological and Experimental Therapeutics, 291,<br />

1017–1022, 1999.<br />

6. Brown, D. M., Wilson, M. R., MacNee, W., Stone, V., and Donaldson, K., Size-dependent proinflammatory<br />

effects of ultrafine polystyrene particles: a role for surface area and oxidative stress in the<br />

enhanced activity of ultrafines, Toxicology and Applied Pharmacology, 175, 191–199, 2001.<br />

7. Donaldson, K. and Tran, C. L., Inflammation caused <strong>by</strong> particles and fibers, Inhalation Toxicology,<br />

14, 5–27, 2002.<br />

8. Gref, R., Luck, M., Quellec, P., Marchand, M., Dellacherie, E., Harnisch, S., Blunk, T., and Muller,<br />

R. H., ‘Stealth’ corona-core nanoparticles surface modified <strong>by</strong> polyethylene glycol (PEG): Influences<br />

of the corona (PEG chain length and surface density) and of the core composition on<br />

phagocytic uptake and plasma protein adsorption, Colloids and Surfaces B, 18, 301–313, 2000.<br />

9. Leroux, J. C., Gravel, P., Balant, L., Volet, B., Anner, B. M., Allemann, E., Doelker, E., and Gurny, R.,<br />

Internalization of poly(D,L-Lactic Acid) nanoparticles <strong>by</strong> isolated human-leukocytes and analysis of<br />

plasma-proteins adsorbed onto the particles, Journal of Biomedical Materials Research, 28, 471–481,<br />

1994.<br />

10. Reynolds, L. J. and Richards, R. J., Can toxicogenomics provide information on the bioreactivity of<br />

diesel exhaust particles? Toxicology, 165, 145–152, 2001.<br />

11. Oberdorster, G., Oberdorster, E., and Oberdorster, J., Nanotoxicology: An emerging discipline<br />

evolving from studies of ultrafine particles, Environmental Health Perspectives, 113, 823–839, 2005.<br />

12. Van de Coevering, R., Kreiter, R., Cardinali, F., van Koten, G., Nierengarten, J. F., and Gebbink, R.,<br />

An octa-cationic core–shell dendrimer as a molecular template for the assembly of anionic fullerene<br />

derivatives, Tetrahedron Letters, 46, 3353–3356, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


130<br />

Nanotechnology for Cancer Therapy<br />

13. Tomalia, D. A., Naylor, A. M., and Goddard, W. A., Starburst dendrimers—molecular-level control<br />

of size shape, surface-chemistry, topology, and flexibility from atoms to macroscopic matter, Angewandte<br />

Chemie-International Edition in English, 29, 138–175, 1990.<br />

14. Tomalia, D. A., Baker, H., Dewald, J., Hall, M., Kallos, G., Martin, S., Roeck, J., Ryder, J., and<br />

Smith, P., A new class of polymers—starburst-dendritic macromolecules, Polymer Journal, 17,<br />

117–132, 1985.<br />

15. Islam, M. T., Majoros, I. J., and Baker, J. R., HPLC analysis of PAMAM dendrimer based multifunctional<br />

devices, Journal of Chromatography B, 822, 21–26, 2005.<br />

16. Binnig, G., Quate, C. F., and Gerber, C., Atomic force microscope, Physical Review Letters, 56,<br />

930–933, 1986.<br />

17. Mecke, A., Lee, D. K., Ramamoorthy, A., Orr, B. G., and Holl, M. M. B., Synthetic and natural<br />

polycationic polymer nanoparticles interact selectively with fluid-phase domains of DMPC lipid<br />

bilayers, Langmuir, 21, 8588–8590, 2005.<br />

18. Mecke, A., Majoros, I. J., Patri, A. K., Baker, J. R., Holl, M. M. B., and Orr, B. G., Lipid bilayer<br />

disruption <strong>by</strong> polycationic polymers: The roles of size and chemical functional group, Langmuir, 21,<br />

10348–10354, 2005.<br />

19. Patri, A. K., Myc, A., Beals, J., Thomas, T. P., Bander, N. H., and Baker, J. R., Synthesis and in vitro<br />

testing of J591 antibody–dendrimer conjugates for targeted prostate cancer therapy, Bioconjugate<br />

Chemistry, 15, 1174–1181, 2004.<br />

20. Malik, N., Wiwattanapatapee, R., Klopsch, R., Lorenz, K., Frey, H., Weener, J. W., Meijer, E. W.,<br />

Paulus, W., and Duncan, R., Dendrimers: Relationship between structure and biocompatibility<br />

in vitro, and preliminary studies on the biodistribution of I-125-labelled polyamidoamine dendrimers<br />

in vivo, Journal of Controlled Release, 65, 133–148, 2000.<br />

21. Hawley, A. E., Davis, S. S., and Illum, L., Targeting of colloids to lymph nodes: Influence of<br />

lymphatic physiology and colloidal characteristics, Advanced Drug Delivery Reviews, 17,<br />

129–148, 1995.<br />

22. Stolnik, S., Illum, L., and Davis, S. S., Long circulating microparticulate drug carriers, Advanced<br />

Drug Delivery Reviews, 16, 195–214, 1995.<br />

23. Ishida, O., Maruyama, K., Sasaki, K., and Iwatsuru, M., Size-dependent extravasation and interstitial<br />

localization of polyethyleneglycol liposomes in solid tumor-bearing mice, International Journal of<br />

Pharmaceutics, 190, 49–56, 1999.<br />

24. Maeda, H., Wu, J., Sawa, T., Matsumura, Y., and Hori, K., Tumor vascular permeability and the EPR<br />

effect in macromolecular therapeutics: A review, Journal of Controlled Release, 65, 271–284, 2000.<br />

25. Brigger, I., Dubernet, C., and Couvreur, P., Nanoparticles in cancer therapy and diagnosis, Advanced<br />

Drug Delivery Reviews, 54, 631–651, 2002.0<br />

26. Kawasaki, E. S. and Player, A., Nanotechnology, nanomedicine, and the development of new,<br />

effective therapies for cancer, Nanomedicine: Nanotechnology, Biology and Medicine, 1,<br />

101–109, 2005.<br />

27. Torchilin, V. P., Drug targeting, European Journal of Pharmaceutical Sciences, 11, S81–S91, 2000.<br />

28. Prosa, T. J., Bauer, B. J., and Amis, E. J., From stars to spheres: A SAXS analysis of dilute dendrimer<br />

solutions, Macromolecules, 34, 4897–4906, 2001.<br />

29. Nisato, G., Ivkov, R., and Amis, E. J., Structure of charged dendrimer solutions as seen <strong>by</strong> smallangle<br />

neutron scattering, Macromolecules, 32, 5895–5900, 1999.<br />

30. Tsay, J. M., Doose, S., and Weiss, S., Rotational and translational diffusion of peptide-coated<br />

CdSe/CdS/ZnS nanorods studied <strong>by</strong> fluorescence correlation spectroscopy, Journal of the American<br />

Chemical Society, 128, 1639–1647, <strong>2006</strong>.<br />

31. Calabretta, M., Jamison, J. A., Falkner, J. C., Liu, Y., Yuhas, B. D., Matthews, K. S., and Colvin,<br />

V. L., Analytical ultracentrifugation for characterizing nanocrystals and their bioconjugates, Nano<br />

Letters, 5, 963–967, 2005.<br />

32. Giddings, J. C., Field-flow fractionation: Analysis of macromolecular, colloidal, and particulate<br />

materials, Science, 260, 1456–1465, 1993.<br />

33. Lockman, P. R., Koziara, J. M., Mumper, R. J., and Allen, D. D., Nanoparticle surface charges alter<br />

blood–brain barrier integrity and permeability, Journal of Drug Targeting, 12, 635–641, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 131<br />

34. Kukowska-Latallo, J. F., Candido, K. A., Cao, Z., Nigavekar, S. S., Majoros, I. J., Thomas, T. P.,<br />

Balogh, L. P., Khan, M. K., and Baker, J. R., Nanoparticle targeting of anticancer drug improves<br />

therapeutic response in animal model of human epithelial cancer, Cancer Research, 65, 5317–5324,<br />

2005.<br />

35. Nel, A., Atmosphere: Enhanced: Air pollution-related illness: Effects of particles, Science, 308,<br />

804–806, 2005.<br />

36. Brown, D. M., Stone, V., Findlay, P., MacNee, W., and Donaldson, K., Increased inflammation and<br />

intracellular calcium caused <strong>by</strong> ultrafine carbon black is independent of transition metals or other<br />

soluble components, Occupational and Environmental Medicine, 57, 685–691, 2000.<br />

37. Derfus, A. M., Chan, W. C. W., and Bhatia, S. N., Probing the cytotoxicity of semiconductor<br />

quantum dots, Nano Letters, 4, 11–18, 2004.<br />

38. Joo, S. H., Feitz, A. J., and Waite, T. D., Oxidative degradation of the carbothioate herbicide, molinate,<br />

using nanoscale zero-valent iron, Environmental Science & Technology, 38, 2242–2247, 2004.<br />

39. Nagaveni, K., Sivalingam, G., Hegde, M. S., and Madras, G., Photocatalytic degradation of organic<br />

compounds over combustion-synthesized nano-TiO 2, Environmental Science & Technology, 38,<br />

1600–1604, 2004.<br />

40. Oberdorster, E., Manufactured nanomaterials (fullerenes, C60) induce oxidative stress in the brain of<br />

juvenile largemouth bass, Environmental Health Perspectives, 112, 1058–1062, 2004.<br />

41. Rancan, F., Rosan, S., Boehm, F., Cantrell, A., Brellreich, M., Schoenberger, H., Hirsch, A., and<br />

Moussa, F., Cytotoxicity and photocytotoxicity of a dendritic C(60) mono-adduct and a malonic acid<br />

C(60) tris-adduct on Jurkat cells, Journal of Photochemistry and Photobiology B, 67, 157–162, 2002.<br />

42. Sayes, C., Fortner, J., Guo, W., Lyon, D., Boyd, A., Ausman, K., Tao, Y., Sitharaman, B., Wilson, L.,<br />

Hughes, J., West, J., and Colvin, V. L., The differential cytotoxicity of water-soluble fullerenes,<br />

Nano Letters, 4, 1881–1887, 2004.<br />

43. Wilson, M. R., Lightbody, J. H., Donaldson, K., Sales, J., and Stone, V., Interactions between<br />

ultrafine particles and transition metals in vivo and in vitro, Toxicology and Applied Pharmacology,<br />

184, 172–179, 2002.<br />

44. Yamakoshi, Y., Umezawa, N., Ryu, A., Arakane, K., Miyata, N., Goda, Y., Masumizu, T., and<br />

Nagano, T., Active oxygen species generated from photoexcited fullerene (C60) as potential<br />

medicines: O2-* versus 1O2, Journal of the American Chemical Society, 125, 12803–12809, 2003.<br />

45. Sonvico, F., Mornet, S., Vasseur, S., Dubernet, C., Jaillard, D., Degrouard, J., Hoebeke, J., Duguet,<br />

E., Colombo, P., and Couvreur, P., Folate-conjugated iron oxide nanoparticles for solid tumor<br />

targeting as potential specific magnetic hyperthermia mediators: synthesis, physicochemical<br />

characterization, and in vitro experiments, Bioconjugate Chemistry, 16, 1181–1188, 2005.<br />

46. Thomas, T. P., Patri, A. K., Myc, A., Myaing, M. T., Ye, J. Y., Norris, T. B., and Baker, J. R., In vitro<br />

targeting of synthesized antibody-conjugated dendrimer nanoparticles, Biomacromolecules, 5,<br />

2269–2274, 2004.<br />

47. Kobayashi, H. and Brechbiel, M. W., Nano-sized MRI contrast agents with dendrimer cores,<br />

Advanced Drug Delivery Reviews, 57, 2271–2286, 2005.<br />

48. Shi, X., Lesniak, W., Islam, M. T., MuNiz, M. C., Balogh, L. P., and Baker, J. J. R., Comprehensive<br />

characterization of surface-functionalized poly(amidoamine) dendrimers with acetamide, hydroxyl,<br />

and carboxyl groups, Colloids and Surfaces A, 272, 139–150, <strong>2006</strong>.<br />

49. Shi, X., Majoros, I. J., Patri, A. K., Bi, X., Islam, M. T., Desai, A., Ganser, T. R., and Baker, J. R.,<br />

Molecular heterogeneity analysis of poly(amidoamine) dendrimer-based mono- and multifunctional<br />

nanodevices <strong>by</strong> capillary electrophoresis, Analyst, 131, 374–381, <strong>2006</strong>.<br />

50. Ferrari, M., Cancer nanotechnology: Opportunities and challenges, Nature Reviews Cancer, 5,<br />

161–171, 2005.<br />

51. Patri, A. K., Majoros, I. J., and Baker, J. R., Dendritic polymer macromolecular carriers for drug<br />

delivery, Current Opinion in Chemical Biology, 6, 466–471, 2002.<br />

52. Portney, N. G. and Ozkan, M., Nano-oncology: drug delivery, imaging, and sensing, Analytical and<br />

Bioanalytical Chemistry, 384, 620–630, <strong>2006</strong>.<br />

53. Patri, A. K., Kukowska-Latallo, J. F., and Baker, J. R., Targeted drug delivery with dendrimers:<br />

Comparison of the release kinetics of covalently conjugated drug and non-covalent drug inclusion<br />

complex, Advanced Drug Delivery Reviews, 57, 2203–2214, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


132<br />

Nanotechnology for Cancer Therapy<br />

54. Gillies, E. R. and Frechet, J. M., pH-responsive copolymer assemblies for controlled release of<br />

doxorubicin, Bioconjugate Chemistry, 16, 361–368, 2005.<br />

55. Brucher, E., Kinetic stabilities of gadolinium(III) chelates used as MRI contrast agents, Topics in<br />

Current Chemistry, 221, 103–122, 2002.<br />

56. Hardman, R., A toxicologic review of quantum dots: Toxicity depends on physicochemical and<br />

environmental factors, Environmental Health Perspectives, 114, 165–172, <strong>2006</strong>.<br />

57. Derfus, A. M., Chan, W. C. W., and Bhatia, S. N., Probing the cytotoxicity of semiconductor<br />

quantum dots, Nano Letters, 4, 11–18, 2004.<br />

58. Dubertret, B., Skourides, P., Norris, D. J., Noireaux, V., Brivanlou, A. H., and Libchaber, A., In vivo<br />

imaging of quantum dots encapsulated in phospholipid micelles, Science, 298, 1759–1762, 2002.<br />

59. Hoshino, A., Hanaki, K., Suzuki, K., and Yamamoto, K., Applications of T-lymphoma labeled with<br />

fluorescent quantum dots to cell tracing markers in mouse body, Biochemical and Biophysical<br />

Research Communications, 314, 46–53, 2004.<br />

60. Voura, E. B., Jaiswal, J. K., Mattoussi, H., and Simon, S. M., Tracking metastatic tumor cell<br />

extravasation with quantum dot nanocrystals and fluorescence emission-scanning microscopy,<br />

Nature Medicine, 10, 993–998, 2004.<br />

61. Ueng, T. H., Kang, J. J., Wang, H. W., Cheng, Y. W., and Chiang, L. Y., Suppression of microsomal<br />

cytochrome P450-dependent monooxygenases and mitochondrial oxidative phosphorylation <strong>by</strong> fullerenol,<br />

a polyhydroxylated fullerene C60, Toxicology Letters, 93, 29–37, 1997.<br />

62. Shcharbin, D., Jokiel, M., Klajnert, B., and Bryszewska, M., Effect of dendrimers on pure acetylcholinesterase<br />

activity and structure, Bioelectrochemistry, 68, 56–59, <strong>2006</strong>.<br />

63. Slikker, W., Andersen, M. E., Bogdanffy, M. S., Bus, J. S., Cohen, S. D., Conolly, R. B., David,<br />

R. M., Doerrer, N. G., Dorman, D. C., Gaylor, D. W., Hattis, D., Rogers, J. M., Setzer, R. W.,<br />

Swenberg, J. A., and Wallace, K., Dose-dependent transitions in mechanisms of toxicity: Case<br />

studies, Toxicology and Applied Pharmacology, 201, 226–294, 2004.<br />

64. Oberdorster, G., Finkelstein, J. N., Johnston, C., Gelein, R., Cox, C., Baggs, R., and Elder, A. C.,<br />

Acute pulmonary effects of ultrafine particles in rats and mice, Research Report (Health Effects<br />

Institute), 5–74, 2000. (disc 75–86)<br />

65. Bazile, D. V., Ropert, C., Huve, P., Verrecchia, T., Marlard, M., Frydman, A., Veillard, M., and<br />

Spenlehauer, G., Body distribution of fully biodegradable [14C]-poly(lactic acid) nanoparticles<br />

coated with albumin after parenteral administration to rats, Biomaterials, 13, 1093–1102, 1992.<br />

66. Cagle, D. W., Kennel, S. J., Mirzadeh, S., Alford, J. M., and Wilson, L. J., In vivo studies of<br />

fullerene-based materials using endohedral metallofullerene radiotracers, Proceedings of the<br />

National Academy of Sciences of the United States of America, 96, 5182–5187, 1999.<br />

67. Ogawara, K., Furumoto, K., Takakura, Y., Hashida, M., Higaki, K., and Kimura, T., Surface hydrophobicity<br />

of particles is not necessarily the most important determinant in their in vivo disposition<br />

after intravenous administration in rats, Journal of Controlled Release, 77, 191–198, 2001.<br />

68. Nigavekar, S. S., Sung, L. Y., Llanes, M., El-Jawahri, A., Lawrence, T. S., Becker, C. W., Balogh, L.,<br />

and Khan, M. K., H-3 dendrimer nanoparticle organ/tumor distribution, Pharmaceutical Research,<br />

21, 476–483, 2004.<br />

69. Ballou, B., Lagerholm, B. C., Ernst, L. A., Bruchez, M. P., and Waggoner, A. S., Noninvasive<br />

imaging of quantum dots in mice, Bioconjugate Chemistry, 15, 79–86, 2004.<br />

70. Gharbi, N., Pressac, M., Hadchouel, M., Szwarc, H., Wilson, S. R., and Moussa, F., [60]Fullerene is a<br />

powerful antioxidant in vivo with no acute or subacute toxicity, Nano Letters, 5, 2578–2585, 2005.<br />

71. Fernandezurrusuno, R., Fattal, E., Porquet, D., Feger, J., and Couvreur, P., Evaluation of liver<br />

toxicological effects induced <strong>by</strong> polyalkylcyanoacrylate nanoparticles, Toxicology and Applied<br />

Pharmacology, 130, 272–279, 1995.<br />

72. Roberts, J. C., Bhalgat, M. K., and Zera, R. T., Preliminary biological evaluation of polyamidoamine<br />

(PAMAM) starburst dendrimers, Journal of Biomedical Materials Research, 30, 53–65, 1996.<br />

73. Oberdorster, G., Maynard, A., Donaldson, K., Castranova, V., Fitzpatrick, J., Ausman, K., Carter, J.,<br />

Karn, B., Kreyling, W., Lai, D., Olin, S., Monteiro-Riviere, N., Warheit, D., and Yang, H., Principles<br />

for characterizing the potential human health effects from exposure to nanomaterials: elements of a<br />

screening strategy, Particle and Fibre Toxicology, 2, 8, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 133<br />

74. Dierickx, P., Prediction of human acute toxicity <strong>by</strong> the hep G2/24-hour/total protein assay, with<br />

protein measurement <strong>by</strong> the CBQCA method, Alternatives to Laboratory Animals, 33, 207–213,<br />

2005.<br />

75. Knasmuller, S., Parzefall, W., Sanyal, R., Ecker, S., Schwab, C., Uhl, M., Mersch-Sundermann, V.,<br />

Williamson, G., Hietsch, G., Langer, T., Darroudi, F., and Natarajan, A. T., Use of metabolically<br />

competent human hepatoma cells for the detection of mutagens and antimutagens, Mutation<br />

Research, 402, 185–202, 1998.<br />

76. Urani, C., Doldi, M., Crippa, S., and Camatini, M., Human-derived cell lines to study xenobiotic<br />

metabolism, Chemosphere, 37, 2785–2795, 1998.<br />

77. Wang, K., Shindoh, H., Inoue, T., and Horii, I., Advantages of in vitro cytotoxicity testing <strong>by</strong> using<br />

primary rat hepatocytes in comparison with established cell lines, Journal of Toxicological Sciences,<br />

27, 229–237, 2002.<br />

78. Gharbi, N., Pressac, M., Tomberli, V., Da Ros, T., Brettreich, M., Hadchouel, M., Arbeille, B.,<br />

Trivin, F., Ceolin, R., Hirsch, A., Prato, M., Szwarc, H., Bensasson, R., and Moussa, F., In Fullerenes<br />

2000: Functionalized Fullerenes, Maggini, M., Martin, N., and Guldi, D. M., Eds., vol. 9, The<br />

Electrochemical Society, Pennington, NJ, 2000.<br />

79. Lee, C. C., MacKay, J. A., Frechet, J. M., and Szoka, F. C., Designing dendrimers for biological<br />

applications, Nature Biotechnology, 23, 1517–1526, 2005.<br />

80. Wang, H., Wang, J., Deng, X., Sun, H., Shi, Z., Gu, Z., Liu, Y., and Zhao, Y., Biodistribution of<br />

carbon single-wall carbon nanotubes in mice, Journal of Nanoscience and Nanotechnology, 4,<br />

1019–1024, 2004.<br />

81. Manil, L., Couvreur, P., and Mahieu, P., Acute renal toxicity of doxorubicin (adriamycin)-loaded<br />

cyanoacrylate nanoparticles, Pharmaceutical Research, 12, 85–87, 1995.<br />

82. Manil, L., Davin, J. C., Duchenne, C., Kubiak, C., Foidart, J., Couvreur, P., and Mahieu, P., Uptake<br />

of nanoparticles <strong>by</strong> rat glomerular mesangial cells in-vivo and in-vitro, Pharmaceutical Research,<br />

11, 1160–1165, 1994.<br />

83. Shiohara, A., Hoshino, A., Hanaki, K., Suzuki, K., and Yamamoto, K., On the cyto-toxicity caused<br />

<strong>by</strong> quantum dots, Microbiology and Immunology, 48, 669–675, 2004.<br />

84. Wang, B., Feng, W. Y., Wang, T. C., Jia, G., Wang, M., Shi, J. W., Zhang, F., Zhao, Y. L., and Chai,<br />

Z. F., Acute toxicity of nano- and micro-scale zinc powder in healthy adult mice, Toxicology Letters,<br />

161, 115–123, <strong>2006</strong>.<br />

85. Toutain, H. and Morin, J. P., Renal proximal tubule cell cultures for studying drug-induced nephrotoxicity<br />

and modulation of phenotype expression <strong>by</strong> medium components, Renal Failure, 14,<br />

371–383, 1992.<br />

86. Mickuviene, I., Kirveliene, V., and Juodka, B., Experimental survey of non-clonogenic viability<br />

assays for adherent cells in vitro, Toxicology In Vitro, 18, 639–648, 2004.<br />

87. Hong, S., Bielinska, A. U., Mecke, A., Keszler, B., Beals, J. L., Shi, X., Balogh, L., Orr, B. G., Baker,<br />

J. R., and Banaszak Holl, M. M., Interaction of poly(amidoamine) dendrimers with supported lipid<br />

bilayers and cells: hole formation and the relation to transport, Bioconjugate Chemistry, 15, 774–782,<br />

2004.<br />

88. Decker, T. and Lohmann-Matthes, M. L., A quick and simple method for the quantitation of lactate<br />

dehydrogenase release in measurements of cellular cytotoxicity and tumor necrosis factor (TNF)<br />

activity, Journal of Immunological Methods, 115, 61–69, 1988.<br />

89. Korzeniewski, C. and Callewaert, D. M., An enzyme-release assay for natural cytotoxicity, Journal<br />

of Immunological Methods, 64, 313–320, 1983.<br />

90. Alley, M. C., Scudiero, D. A., Monks, A., Hursey, M. L., Czerwinski, M. J., Fine, D. L., Abbott, B. J.,<br />

Mayo, J. G., Shoemaker, R. H., and Boyd, M. R., Feasibility of drug screening with panels of human<br />

tumor cell lines using a microculture tetrazolium assay, Cancer Research, 48, 589–601, 1988.<br />

91. Berridge, M. V., Herst, P. M., and Tan, A. S., Tetrazolium dyes as tools in cell biology: New insights<br />

into their cellular reduction, Biotechnology Annual Review, 11, 127–152, 2005.<br />

92. Natarajan, M., Mohan, S., Martinez, B. R., Meltz, M. L., and Herman, T. S., Antioxidant compounds<br />

interfere with the 3, Cancer Detection and Prevention, 24, 405–414, 2000.<br />

93. Vellonen, K. S., Honkakoski, P., and Urtti, A., Substrates and inhibitors of efflux proteins interfere<br />

with the MTT assay in cells and may lead to underestimation of drug toxicity, European Journal of<br />

Pharmaceutical Sciences, 23, 181–188, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


134<br />

Nanotechnology for Cancer Therapy<br />

94. Voigt, W., Sulforhodamine B assay and chemosensitivity, Methods in Molecular Medicine, 110,<br />

39–48, 2005.<br />

95. Tuschl, H. and Schwab, C. E., Flow cytometric methods used as screening tests for basal toxicity of<br />

chemicals, Toxicology In Vitro, 18, 483–491, 2004.<br />

96. Cui, D., Tian, F., Ozkan, C. S., Wang, M., and Gao, H., Effect of single wall carbon nanotubes on<br />

human HEK293 cells, Toxicology Letters, 155, 73–85, 2005.<br />

97. Tao, F., Gonzalez-Flecha, B., and Kobzik, L., Reactive oxygen species in pulmonary inflammation<br />

<strong>by</strong> ambient particulates, Free Radical Biology and Medicine, 35, 327–340, 2003.<br />

98. Li, N., Sioutas, C., Cho, A., Schmitz, D., Misra, C., Sempf, J., Wang, M., Oberley, T., Froines, J., and<br />

Nel, A., Ultrafine particulate pollutants induce oxidative stress and mitochondrial damage, Environmental<br />

Health Perspectives, 111, 455–460, 2003.<br />

99. Fernandez-Urrusuno, R., Fattal, E., Rodrigues, J. M., Feger, J., Bedossa, P., and Couvreur, P., Effect<br />

of polymeric nanoparticle administration on the clearance activity of the mononuclear phagocyte<br />

system in mice, Journal of Biomedical Materials Research, 31, 401–408, 1996.<br />

100. Fernandez-Urrusuno, R., Fattal, E., Feger, J., Couvreur, P., and Therond, P., Evaluation of hepatic<br />

antioxidant systems after intravenous administration of polymeric nanoparticles, Biomaterials, 18,<br />

511–517, 1997.<br />

101. Lovric, J., Cho, S. J., Winnik, F. M., and Maysinger, D., Unmodified cadmium telluride quantum<br />

dots induce reactive oxygen species formation leading to multiple organelle damage and cell death,<br />

Chemical Biology, 12, 1227–1234, 2005.<br />

102. Risom, L., Moller, P., and Loft, S., Oxidative stress-induced DNA damage <strong>by</strong> particulate air pollution,<br />

Mutation Research, 592, 119–137, 2005.<br />

103. Grisham, M. B. and McCord, J. M., Chemistry and cytotoxicity of reactive oxygen metabolites, In<br />

Physiology of Oxygen Radicals, <strong>Taylor</strong>, A. E., Matalon, S., and Ward, P., Eds., American Physiology<br />

Society, Bethesda, MD, pp. 1–18, 1986.<br />

104. Kamat, J. P., Devasagayam, T. P. A., Priyadarsini, K. I., and Mohan, H., Reactive oxygen species<br />

mediated membrane damage induced <strong>by</strong> fullerene derivatives and its possible biological implications,<br />

Toxicology, 155, 55–61, 2000.<br />

105. Dotan, Y., Lichtenberg, D., and Pinchuk, I., Lipid peroxidation cannot be used as a universal<br />

criterion of oxidative stress, Progress in Lipid Research, 43, 200–227, 2004.<br />

106. Black, M. J. and Brandt, R. B., Spectrofluorometric analysis of hydrogen peroxide, Analytical<br />

Biochemistry, 58, 246–254, 1974.<br />

107. Dubuisson, M. L., de Wergifosse, B., Trouet, A., Baguet, F., Marchand-Brynaert, J., and Rees, J. F.,<br />

Antioxidative properties of natural coelenterazine and synthetic methyl coelenterazine in rat hepatocytes<br />

subjected to tert-butyl hydroperoxide-induced oxidative stress, Biochemical Pharmacology,<br />

60, 471–478, 2000.<br />

108. Shaik, I. H. and Mehvar, R., Rapid determination of reduced and oxidized glutathione levels using a<br />

new thiol-masking reagent and the enzymatic recycling method: application to the rat liver and bile<br />

samples, Analytical and Bioanalytical Chemistry, 385 (1), 105–113, <strong>2006</strong>.<br />

109. Van Cruchten, S. and Van den Broeck, W., Morphological and biochemical aspects of apoptosis,<br />

oncosis and necrosis, Anatomia, Histologia, Embryologia: Journal of Veterinary Medicine Series C,<br />

31, 214–223, 2002.<br />

110. Bottini, M., Bruckner, S., Nika, K., Bottini, N., Bellucci, S., Magrini, A., Bergamaschi, A., and<br />

Mustelin, T., Multi-walled carbon nanotubes induce T lymphocyte apoptosis, Toxicology Letters,<br />

160, 121–126, <strong>2006</strong>.<br />

111. Cui, D., Tian, F., Ozkan, C. S., Wang, M., and Gao, H., Effect of single wall carbon nanotubes on<br />

human HEK293 cells, Toxicology Letters, 155, 73–85, 2005.<br />

112. Kuo, J. H., Jan, M. S., and Chiu, H. W., Mechanism of cell death induced <strong>by</strong> cationic dendrimers in RAW<br />

264.7 murine macrophage-like cells, Journal of Pharmacy and Pharmacology, 57, 489–495, 2005.<br />

113. Cui, D., Tian, F., Ozkan, C. S., Wang, M., and Gao, H., Effect of single wall carbon nanotubes on<br />

human HEK293 cells, Toxicology Letters, 155, 73–85, 2005.<br />

114. Wang, Z. B., Liu, Y. Q., and Cui, Y. F., Pathways to caspase activation, Cell Biology International,<br />

29, 489–496, 2005.<br />

115. Gurtu, V., Kain, S. R., and Zhang, G., Fluorometric and colorimetric detection of caspase activity<br />

associated with apoptosis, Analytical Biochemistry, 251, 98–102, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 135<br />

116. Loo, D. T. and Rillema, J. R., Measurement of cell death, Methods in Cell Biology, 57, 251–264, 1998.<br />

117. Le Bras, M., Clement, M. V., Pervaiz, S., and Brenner, C., Reactive oxygen species and the mitochondrial<br />

signaling pathway of cell death, Histology and Histopathology, 20, 205–219, 2005.<br />

118. Xia, T., Korge, P., Weiss, J. N., Li, N., Venkatesen, M. I., Sioutas, C., and Nel, A., Quinones and<br />

aromatic chemical compounds in particulate matter induce mitochondrial dysfunction: implications<br />

for ultrafine particle toxicity, Environmental Health Perspectives, 112, 1347–1358, 2004.<br />

119. Hussain, S. M., Hess, K. L., Gearhart, J. M., Geiss, K. T., and Schlager, J. J., In vitro toxicity of<br />

nanoparticles in BRL 3A rat liver cells, Toxicology In Vitro, 19, 975–983, 2005.<br />

120. Foley, S., Crowley, C., Smaihi, M., Bonfils, C., Erlanger, B. F., Seta, P., and Larroque, C., Cellular<br />

localisation of a water-soluble fullerene derivative, Biochemical and Biophysical Research Communications,<br />

294, 116–119, 2002.<br />

121. Isakovic, A., Markovic, Z., Todorovic-Markovic, B., Nikolic, N., Vranjes-Djuric, S., Mirkovic, M.,<br />

Dramicanin, M., Harhaji, L., Raicevic, N., Nikolic, Z., and Trajkovic, V., Distinct cytotoxic<br />

mechanisms of pristine versus hydroxylated fullerene, Toxicological Sciences, 91 (1), 173–183, <strong>2006</strong>.<br />

122. Qi, L. F., Xu, Z. R., Li, Y., Jiang, X., and Han, X. Y., In vitro effects of chitosan nanoparticles on<br />

proliferation of human gastric carcinoma cell line MGC803 cells, World Journal of Gastroenterology,<br />

11, 5136–5141, 2005.<br />

123. Hirsch, T., Susin, S. A., Marzo, I., Marchetti, P., Zamzami, N., and Kroemer, G., Mitochondrial<br />

permeability transition in apoptosis and necrosis, Cell Biology and Toxicology, 14, 141–145, 1998.<br />

124. Amacher, D. E., Drug-associated mitochondrial toxicity and its detection, Current Medicinal Chemistry,<br />

12, 1829–1839, 2005.<br />

125. Gogvadze, V., Orrenius, S., and Zhivotovsky, B., Analysis of mitochondrial dysfunction during cell<br />

death. In Current Protocols in Toxicology, Wiley, New York, pp. 10.1–2.10.27, 2004.<br />

126. Guo, W.-X., Pye, Q. N., Williamson, K. S., Stewart, C. A., Hensley, K. L., Kotake, Y., Floyd, R. A.,<br />

and Broyles, R. H., Mitochondrial dysfunction in choline deficiency-induced apoptosis in cultured<br />

rat hepatocytes, Free Radical Biology and Medicine, 39, 641–650, 2005.<br />

127. Merrick, B. A. and Madenspacher, J. H., Complementary gene and protein expression studies and<br />

integrative approaches in toxicogenomics, Toxicology and Applied Pharmacology, 207, 189–194, 2005.<br />

128. Jani, P., Halbert, G. W., Langridge, J., and Florence, A. T., Nanoparticle uptake <strong>by</strong> the rat gastrointestinal<br />

mucosa: Quantitation and particle size dependency, Journal of Pharmacy and Pharmacology, 42,<br />

821–826, 1990.<br />

129. Duncan, R., Polymer conjugates for tumour targeting and intracytoplasmic delivery. The EPR effect<br />

as a common gateway? Pharmaceutical Science & Technology Today, 2, 441–449, 1999.<br />

130. Briley-Saebo, K., Hustvedt, S. O., Haldorsen, A., and Bjornerud, A., Long-term imaging effects in<br />

rat liver after a single injection of an iron oxide nanoparticle based MR contrast agent, Journal of<br />

Magnetic Resonance Imaging, 20, 622–631, 2004.<br />

131. Cahouet, A., Denizot, B., Hindre, F., Passirani, C., Heurtault, B., Moreau, M., Le Jeune, J., and<br />

Benoit, J., Biodistribution of dual radiolabeled lipidic nanocapsules in the rat using scintigraphy and<br />

gamma counting, International Journal of Pharmaceutics, 242, 367–371, 2002.<br />

132. Deen, W. M., Lazzara, M. J., and Myers, B. D., Structural determinants of glomerular permeability,<br />

American Journal of Physiology—Renal Physiology, 281, F579–F596, 2001.<br />

133. Walton, K., Dorne, J. L., and Renwick, A. G., Species-specific uncertainty factors for compounds<br />

eliminated principally <strong>by</strong> renal excretion in humans, Food and Chemical Toxicology, 42, 261–274, 2004.<br />

134. Olson, H., Betton, G., Robinson, D., Thomas, K., Monro, A., Kolaja, G., Lilly, P., Sanders, J., Sipes,<br />

G., Bracken, W., Dorato, M., Van Deun, K., Smith, P., Berger, B., and Heller, A., Concordance of the<br />

toxicity of pharmaceuticals in humans and in animals, Regulatory Toxicology and Pharmacology,<br />

32, 56–67, 2000.<br />

135. Neerman, M. F., Zhang, W., Parrish, A. R., and Simanek, E. E., In vitro and in vivo evaluation of a<br />

melamine dendrimer as a vehicle for drug delivery, International Journal of Pharmaceutics, 281,<br />

129–132, 2004.<br />

136. Harries, M., Ellis, P., and Harper, P., Nanoparticle albumin-bound paclitaxel for metastatic breast<br />

cancer, Journal of Clinical Oncology, 23, 7768–7771, 2005.<br />

137. Descotes, J., Importance of immunotoxicity in safety assessment: a medical toxicologist’s perspective,<br />

Toxicology Letters, 149, 103–108, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


136<br />

Nanotechnology for Cancer Therapy<br />

138. Dobrovolskaia, M. A. and Kozlov, S. V., Inflammation and cancer: When NF-kappa B amalgamates<br />

the perilous partnership, Current Cancer Drug Targets, 5, 325–344, 2005.<br />

139. Merk, H. F., Sachs, B., and Baron, J., The skin: Target organ in immunotoxicology of smallmolecular-weight<br />

compounds, Skin Pharmacology and Applied Skin Physiology, 14, 419–430, 2001.<br />

140. Myers, A., Clark, J., and Foster, H., Tuberculosis and treatment with infliximab, New England<br />

Journal of Medicine, 346, 625–626, 2002.<br />

141. Keane, J., Gershon, S., Wise, R. P., Mirabile-Levens, E., Kasznica, J., Schwieterman, W. D., Siegel,<br />

J. N., and Braun, M. M., Tuberculosis associated with infliximab, a tumor necrosis factor (alpha)neutralizing<br />

agent, New England Journal of Medicine, 345, 1098–1104, 2001.<br />

142. Zhang, Z., Correa, H., and Begue, R. E., Tuberculosis and treatment with infliximab, New England<br />

Journal of Medicine, 346, 623–626, 2002.<br />

143. Lim, W. S., Powell, R. J., and Johnston, I. D., Tuberculosis and treatment with infliximab, New<br />

England Journal Medicine, 346, 623–626, 2002.<br />

144. Nart, D., Nalbantgil, S., Yagdi, T., Yilmaz, F., Hekimgil, M., Yuce, G., and Hamulu, A., Primary<br />

cardiac lymphoma in a heart transplant recipient, Transplant Proceedings, 37, 1362–1364, 2005.<br />

145. Caillard, S., Pencreach, E., Braun, L., Marcellin, L., Jaegle, M. L., Wolf, P., Parissiadis, A.,<br />

Hannedouche, T., Gaub, M. P., and Moulin, B., Simultaneous development of lymphoma in recipients<br />

of renal transplants from a single donor: Donor origin confirmed <strong>by</strong> human leukocyte antigen<br />

staining and microsatellite analysis, Transplantation, 79, 79–84, 2005.<br />

146. Karakus, S., Ozyilkan, O., Akcali, Z., Demirhan, B., and Haberal, M., Acute myeloid leukemia 4<br />

years after Kaposi’s sarcoma in a renal transplant recipient, Onkologie, 27, 163–165, 2004.<br />

147. http://ncl.cancer.gov/working_assay-cascade.asp<br />

148. FDA/CDER, Guidance for industry. Immunotoxicology evaluation of investigational new drugs, 2002,<br />

http://www.fda.gov/Cder/guidance/.<br />

149. FDA/CDER, ICH S8: Immunotoxicity studies for human pharmaceuticals (draft), 2004, http://www.<br />

fda.gov/Cder/guidance/.<br />

150. FDA/CBER/CDER, Guidance for industry. Developing medical imaging drug and biological<br />

products. Part 1: Conducting safety assessments, 2004, http://www.fda.gov/Cder/guidance/.<br />

151. FDA/CDER/CBER, Guidance for industry. ICH S6: Preclinical safety evaluation of biotechnologyderived<br />

pharmaceuticals, 1997, http://www.fda.gov/Cder/guidance/.<br />

152. FDA/CDRH, Guidance for industry and FDA reviewers. Immunotoxicity testing guidance, May<br />

1999, http://www.fda.gov/cdrh/guidance.html.<br />

153. ANSI/AAMI/ISO, 10993-4: Biological evaluation of medical devices. Part 4: Selection of tests for<br />

interaction with blood, 2002, http://www.iso.org/iso/en/CatalogueListPage.CatalogueList.<br />

154. ASTM, Standard practice. F1906-98: Evaluation of immune responses in biocompatibility testing<br />

using ELISA tests, lymphocyte proliferation, and cell migration, 2003, http://www.astm.org/cgibin/<br />

SoftCart.exe/NEWSITE_JAVASCRIPT/index.shtml?L+mystore+zdrn2970+1157995562.<br />

155. ASTM, Standard practice. F748-98: Selecting generic biological test methods for<br />

materials and devices, http://www.astm.org/cgibin/SoftCart.exe/NEWSITE_JAVASCRIPT/index.<br />

shtml?L+mystore+zdrn2970+1157995562.<br />

156. ASTM, F756-00: Standard practice for assessment of hemolytic properties of materials, 2000,<br />

http://www.astm.org/cgibin/SoftCart.exe/NEWSITE_JAVASCRIPT/index.shtml?L+mystore+zdrn<br />

2970+1157995562.<br />

157. Balakrishnan, B., Kumar, D. S., Yoshida, Y., and Jayakrishnan, A., Chemical modification of<br />

poly(vinyl chloride) resin using poly(ethylene glycol) to improve blood compatibility, Biomaterials,<br />

26, 3495–3502, 2005.<br />

158. Oyewumi, M. O., Yokel, R. A., Jay, M., Coakley, T., and Mumper, R. J., Comparison of cell uptake,<br />

biodistribution and tumor retention of folate-coated and PEG-coated gadolinium nanoparticles in<br />

tumor-bearing mice, Journal of Controlled Release, 95, 613–626, 2004.<br />

159. Diederichs, J. E., Plasma protein adsorption patterns on liposomes: establishment of analytical<br />

procedure, Electrophoresis, 17, 607–611, 1996.<br />

160. Harnisch, S. and Muller, R. H., Plasma protein adsorption patterns on emulsions for parenteral<br />

administration: establishment of a protocol for two-dimensional polyacrylamide electrophoresis,<br />

Electrophoresis, 19, 349–354, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Preclinical Characterization of Engineered Nanoparticles Intended for Cancer Therapeutics 137<br />

161. Goppert, T. M. and Muller, R. H., Alternative sample preparation prior to two-dimensional electrophoresis<br />

protein analysis on solid lipid nanoparticles, Electrophoresis, 25, 134–140, 2004.<br />

162. Thode, K., Luck, M., Semmler, W., Muller, R. H., and Kresse, M., Determination of plasma protein<br />

adsorption on magnetic iron oxides: sample preparation, Pharmaceutial Research, 14, 905–910, 1997.<br />

163. Peracchia, M. T., Harnisch, S., Pinto-Alphandary, H., Gulik, A., Dedieu, J. C., Desmaele, D.,<br />

d’Angelo, J., Muller, R. H., and Couvreur, P., Visualization of in vitro protein-rejecting properties<br />

of PEGylated stealth (R) polycyanoacrylate nanoparticles, Biomaterials, 20, 1269–1275, 1999.<br />

164. Gessner, A., Lieske, A., Paulke, B. R., and Muller, R. H., Influence of surface charge density on<br />

protein adsorption on polymeric nanoparticles: Analysis <strong>by</strong> two-dimensional electrophoresis,<br />

European Journal of Pharmaceutics and Biopharmaceutics, 54, 165–170, 2002.<br />

165. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

Theory to practice, Pharmacological Reviews, 53, 283–318, 2001.<br />

166. Blunk, T., Hochstrasser, D. F., Sanchez, J. C., Muller, B. W., and Muller, R. H., Colloidal carriers for<br />

intravenous drug targeting: Plasma protein adsorption patterns on surface-modified latex particles<br />

evaluated <strong>by</strong> two-dimensional polyacrylamide gel electrophoresis, Electrophoresis, 14, 1382–1387,<br />

1993.<br />

167. Diluzio, N. R. and Zilversmit, D. B., Influence of exogenous proteins on blood clearance and tissue<br />

distribution of colloidal gold, American Journal of Physiology, 180, 563–565, 1955.<br />

168. Campbell, P. A., Canono, B. P., and Drevets, D. A., Measurement of bacterial ingestion and killing<br />

<strong>by</strong> macrophages, Current Protocols in Immunology, Suppl. 12, 14.6.1–14.6.13, 1994.<br />

169. Prabha, S., Zhou, W. Z., Panyam, J., and Labhasetwar, V., Size-dependency of nanoparticlemediated<br />

gene transfection: studies with fractionated nanoparticles, International Journal of Pharmaceutics,<br />

244, 105–115, 2002.<br />

170. Chamberlain, P. and Mire-Sluis, A. R., An overview of scientific and regulatory issues for the<br />

immunogenicity of biological products, Developments in Biological Standardization (Basel), 112,<br />

3–11, 2003.<br />

171. Casadevall, N., Antibodies against rHuEPO: native and recombinant, Nephrology Dialysis Transplantation,<br />

17(Suppl. 5), 42–47, 2002.<br />

172. Casadevall, N., Nataf, J., Viron, B., Kolta, A., Kiladjian, J. J., Martin-Dupont, P., Michaud, P., Papo,<br />

T., Ugo, V., Teyssandier, I., Varet, B., and Mayeux, P., Pure red-cell aplasia and antierythropoietin<br />

antibodies in patients treated with recombinant erythropoietin, New England Journal of Medicine,<br />

346, 469–475, 2002.<br />

173. Koren, E., Zuckerman, L. A., and Mire-Sluis, A. R., Immune responses to therapeutic proteins in<br />

humans—clinical significance, assessment and prediction, Current Pharmaceutical Biotechnology,3,<br />

349–360, 2002.<br />

174. Swanson, S. J., Ferbas, J., Mayeux, P., and Casadevall, N., Evaluation of methods to detect and<br />

characterize antibodies against recombinant human erythropoietin, Nephron Clinical Practice, 96,<br />

c88–c95, 2004.<br />

175. Andreev, S. M., Babakhin, A. A., Petrukhina, A. O., Romanova, V. S., Parnes, Z. N., and Petrov,<br />

R. V., Immunogenetic and allergenic activities conjugates of fullere with amino acids and protein,<br />

Doklady Akademii Nauk, 370, 261–264, 2000.<br />

176. Braden, B. C., Goldbaum, F. A., Chen, B. X., Kirschner, A. N., Wilson, S. R., and Erlanger, B. F., Xray<br />

crystal structure of an anti-Buckminsterfullerene antibody Fab fragment: Biomolecular recognition<br />

of C-60, Proceedings of the National Academy of Sciences of the United States of America, 97,<br />

12193–12197, 2000.<br />

177. Chen, B. X., Wilson, S. R., Das, M., Coughlin, D. J., and Erlanger, B. F., Antigenicity of fullerenes:<br />

antibodies specific for fullerenes and their characteristics, Proceedings of the National Academy of<br />

Sciences of the United States of America, 95, 10809–10813, 1998.<br />

178. Masalova, O. V., Shepelev, A. V., Atanadze, S. N., Parnes, Z. N., Romanova, V. S., Vol’pina, O. M.,<br />

Semiletov Iu, A., and Kushch, A. A., Immunostimulating effect of water-soluble fullerene derivatives—perspective<br />

adjuvants for a new generation of vaccine, Doklady Akademii Nauk, 369,<br />

411–413, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


8<br />

CONTENTS<br />

Nanotechnology: Regulatory<br />

Perspective for Drug<br />

Development in Cancer<br />

Therapeutics<br />

N. Sadrieh and T. J. Miller<br />

8.1 Introduction ......................................................................................................................... 139<br />

8.2 FDA Experience with Nanotechnology Product Applications........................................... 140<br />

8.3 FDA Definition of Nanotechnology ................................................................................... 143<br />

8.4 FDA Initiatives in Nanotechnology .................................................................................... 143<br />

8.5 Research at the FDA on Nanotechnology .......................................................................... 145<br />

8.6 Regulatory Considerations for Nanotechnology Products ................................................. 145<br />

8.6.1 Jurisdiction of Nanotechnology Products ............................................................... 145<br />

8.6.2 Existing FDA Regulations That Apply to Nanomaterial-Containing Products..... 146<br />

8.7 Scientific Considerations for the Development of Nanotechnology Products .................. 147<br />

8.7.1 Characterization of Nanoparticles........................................................................... 147<br />

8.7.2 Safety Considerations.............................................................................................. 149<br />

8.8 Conclusions ......................................................................................................................... 149<br />

8.9 Trademark Listing ............................................................................................................... 151<br />

References .................................................................................................................................... 151<br />

8.1 INTRODUCTION<br />

In the last decade, remarkable scientific progress in the development of novel, nanoscale technologies<br />

has generated rising expectations for this enabling technology to advance our efforts in the<br />

prevention, diagnosis and treatment of human disease. These nanotechnologies, commonly referred<br />

to as nanomedicines when applied to the treatment, diagnosis, monitoring, and control of biological<br />

systems, incorporate a multitude of diverse and often unique nanosized products including new<br />

and/or improved therapeutic drugs, diagnostic/surgical devices, and targeted drug delivery<br />

systems. 1 The list of nanoparticle medicines approved and/or currently being developed is quite<br />

large; these products range from early forms of drug-encapsulating liposomes (e.g., liposomes<br />

containing doxorubicin for treatment of breast and ovarian cancer) and simple polymeric structures<br />

(e.g., actinomycin D adsorbed on poly(methylcryanoacrylate) (PMCA) and poly(ethylcryanoacrylate)<br />

(PECA) nanosized particles) 2 to more intricate multifunctional “nanoplatforms” and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

139


140<br />

Nanotechnology for Cancer Therapy<br />

diagnostic devices, including polymeric dendrimers, nanocrystal quantum dots, carbon-based fullerenes,<br />

and nanoshells, to name a few (see Table 8.1). The vast majority of these products are being<br />

developed to improve early detection of cancer and/or to provide alternative clinical therapeutic<br />

approaches to fighting cancer, some of these products have been approved <strong>by</strong> the Food and Drug<br />

Administration for clinical use. 75<br />

Although there have been many articles published on the in vitro toxicity of certain nanomaterials,<br />

very little in vivo safety data evaluating the general toxicity of these novel products is<br />

available in the published literature. Many in vitro assays (cell-based, organelle-based, and cell/organelle-free<br />

systems) have been used to evaluate the cytotoxicity of nanomaterials to help predict<br />

the relative safety and general biocompatibility of the nanoproducts in vivo. Several such studies<br />

include viability assays in vascular endothelial cells exposed to C 60; 76 fullerene induction/inhibition<br />

of oxidative stress and lipid peroxidation in isolated microsomes, dermal fibroblasts, and<br />

cultured cortical neurons; 77–79 cytotoxicity of metal oxide nanoparticles to BRL 3A liver cells; 80<br />

and macrophage apoptosis after exposure to carbon nanomaterials and ultrafine particles. 81–83<br />

Although a vast amount of useful information has been obtained from these various in vitro<br />

models, the relevance and application of these findings to human drug safety remain uncertain<br />

until validated with in vivo studies with identical nanomaterials.<br />

Targeted in vivo efficacy data is readily available for many of the novel, potential nanotherapeutics;<br />

however, necessary in vivo data indicating mechanism of action, general<br />

pharmacology/toxicity profiles, and optimal product characterization efforts rarely appear in the<br />

published literature. Much of our current understanding of the biocompatibility of nanomaterials<br />

comes from focused inhalation and dermal studies using nanosized products and ultrafine contaminants,<br />

as a model of environmental and occupational risk. 84–87 However, there are questions<br />

regarding the level of applicability of these studies to human drug safety assessment due to the<br />

variability in size and purity of the chemical or structural composition reported or not reported in<br />

these studies, the route of primary exposure, and the complex differences in the structure and<br />

function of the affected organs. 88 In addition, other in vivo studies, which identify intrinsic antioxidant<br />

and oxidative stress-inducing properties of unsubstituted fullerenes in a rodent and fish<br />

species, respectively, have important but limited value in drug discovery in the absence of systemic,<br />

whole animal toxicology data 67,89 .<br />

While the enthusiasm and expectations for success in utilizing nanomaterials and nanometersized<br />

structures in biomedical applications grow, so do the concerns regarding the potential impact<br />

of such products on the environment and human health. 88 The absence of available safety information<br />

for a vast majority of these novel nanosized materials and nanoparticlized drugs in<br />

biological systems only further supports the need for additional research into any potential toxic<br />

mechanisms in humans. The questions that appear most often in deciding the safety of any nanoscale<br />

material include:<br />

† Are nanomaterials and nanotechnology new to the United States Food and Drug<br />

Administration (FDA)?<br />

† How do existing regulations ensure the development of safe and effective nanotechnology-based<br />

drugs and how is the FDA preparing to deal with nanotechnology?<br />

† What are the scientific considerations that raise unique concerns for nanotechnologybased<br />

therapeutics?<br />

8.2 FDA EXPERIENCE WITH NANOTECHNOLOGY PRODUCT APPLICATIONS<br />

Historically, the FDA has encountered and approved many products with particulate materials<br />

characteristic of a nanomedicine or nanoproduct (e.g., small size, mechanism of delivery, increased<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


TABLE 8.1<br />

Nanomedicines: Drugs/Devices FDA Approved and in Development<br />

Name Structure Company Size (nm) Phase Indication Reference a,b<br />

Magnevist wc<br />

PIO Shering AG !1 APP MRI tumor imaging 3–5<br />

Feridex wc<br />

SPIO Adv.Magnetics 120–180 APP MRI tumor imaging 5–9<br />

Rapamune wc<br />

Nanocrystal drug Wyeth 100–1,000 APP Immunosuppressant agent 10–12<br />

Emend wc<br />

Nanocrystal drug Merck 100–1,000 APP Nausea 13,14<br />

Doxil wc<br />

Liposome-drug encapsulation Ortho Biotech z100 APP Metastatic ovarian cancer 15–17<br />

Abraxane wc<br />

Nanoparticulate albumin American Pharmaceutical z130 APP Non-small-cell lung cancer, breast cancer, 18–20<br />

Partners<br />

others<br />

SilvaGard w<br />

Silver nanoparticles AcryMed z10 APP Antimicrobial surface treatment (medical<br />

devices)<br />

21–23<br />

AmBisome wc<br />

Liposome-drug encapsulation Gilead Science 45–80 APP Fungal infections 24–26<br />

Leunesse w<br />

Solid lipid nanoparticles Aano-therapeutics N/A On market Cosmetics N/A<br />

NB-001 d<br />

Nanoemulsion NanoBio Corp z150 Phase II Topical microbicide for herpes labialis 27–30<br />

VivaGel w<br />

Dendrimer Starpharma Holdings Ltd. N/A Phase I Vaginal microbicide (STI and HIV<br />

prevention)<br />

31–34<br />

INGN-401 c<br />

Drug–liposome complex Introgen Therapeutics N/A Phase I Metastatic lung cancer 35–38<br />

Combidex wc<br />

USPIO Advanced Magnetics 20–50 NDA filed MRI tumor imaging 39–42<br />

IT-101 c<br />

Drug–polymer complex Insert Therapeutics N/A IND filed Metastatic solid tumors 43,44<br />

Dendrimers c<br />

Polymeric spheres, multifunctional Dendritech Nanotech., O1 to!500 IDV Diagnostic, contrast agents, microbicide, 45–54<br />

Starpharma<br />

drug delivery, B neutron capture<br />

Nanoshells c<br />

Metal shell, SiO2 core Nanospectra 100–200 IDV Photo-thermal ablation therapy /imaging<br />

tumors<br />

55–58<br />

Fullerenes c<br />

Carbon sphere, C60, multifunctional C Sixty 1 IDV MRI contrast agent, antiviral, antioxidant 59–67<br />

Quantum dots c<br />

CdSe a nanocrystal Evident Technologies 2–10 IDV Optical imaging, tumor imaging,<br />

photosensitizer<br />

68–74<br />

APP, approved; IDV, in development; PIO, paramagnetic iron oxide; SPIO, superparamagnetic iron oxide; USPIO, ultrasmall superparamagnetic iron oxide; STI, sexually transmitted<br />

infection; HIV, human immunodeficiency virus; N/A, not available.<br />

a<br />

Most typical semiconductor metal used, however can also be composed of nearly any other semiconductor metal (e.g., CdS, CdTe, ZnS, PbS, etc.).<br />

b<br />

References do not necessarily refer to the product described, and may reflect general properties about the product.<br />

c<br />

Drugs with proposed or proven diagnostic or therapeutic benefit to cancer patients.<br />

d<br />

Unable to locate peer-reviewed publications. References listed are press releases found during internet search.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Regulatory Perspective in Nanotechnology Products 141


142<br />

Nanotechnology for Cancer Therapy<br />

surface area, specialized function related to size and increased surface area, etc.). Although the field<br />

of nanomedicine may be new, the submission of applications for products containing novel dosage<br />

forms or drug delivery systems, including nanomaterials, nanoparticlized drugs and medical<br />

devices, is not particularly new to this agency. Products are reviewed on a product-<strong>by</strong>-product<br />

basis, and as such, various concepts of risk management (i.e., risk identification, risk analysis, risk<br />

control, etc.) are often implemented to support the drug review process.<br />

The FDA is aware of several FDA regulated products that employ nanotechnology. However,<br />

to date, few manufacturers of regulated products have claimed the use of nanotechnology in the<br />

manufacture of their products or made any nanotechnology claims for the finished product.<br />

The FDA is aware that a few cosmetic products and sunscreens claim to contain nanoparticles<br />

to increase the product stability, modify release of ingredients, or change the opacity of the<br />

product <strong>by</strong> using nanoparticles of titanium dioxide and zinc oxide. Similarly, several drugs<br />

such as Gd-DTPA (Magnevist w , tumor contrast agent), 3–5 ferumoxides (Feridex w , tumor contrast<br />

agent), 5–9 liposomal doxorubicin (Doxil w , chemotherapeutic), 15–17 and albumin-bound paclitaxel<br />

(Abraxane w , chemotherapeutic), 18–20 to name a few, have all been approved for human clinical<br />

use in the diagnosis and treatment of a variety of patient tumors and metastatic cancers. For public<br />

information regarding FDA approval letters, label information, and review packages for each<br />

of these drugs and others listed in Table 8.1, please visit the following website: http://www.<br />

accessdata.fda.gov/scripts/cder/drugsatfda/.<br />

Of the many drugs listed in Table 8.1, tumor contrast agents comprise some of the earliest<br />

nanosized products to receive FDA approval for human clinical use. Critical to therapeutic<br />

outcome, diagnostic imaging of nanosized contrast agents using non-invasive magnetic resonance<br />

imaging (MRI) and positron emission tomography (PET), have shown advanced utility for early<br />

detection of small tumors, metastatic lymph nodes, and altered tumor microenvironments. 90<br />

The leading MR contrast agent, Magnevist w , a paramagnetic, gadolinium-based contrast<br />

medium, and Feridex w , a superparamagnetic iron oxide contrast agent, received primary FDA<br />

approval for clinical use in 1988 and 1996, respectively. Each contrast agent has a different<br />

mean particle size (!1 nm and 120–180 nm, respectively), is approved for different clinical indications,<br />

and demonstrates product-specific adverse sensitivity and pharmacology/toxicology<br />

profiles in patients. 3–9<br />

Another early nanomedicine, liposomal encapsulated doxorubicin (Doxil w ), is regulated <strong>by</strong> the<br />

FDA and has been available in the clinic for treatment of various cancers since 1995. Drug<br />

incorporation inside the hydrophilic core or within the hydrophobic phospholipid bilayer coat of<br />

liposomes, has been shown to improve drug solubility, enhance drug transfer into cells and tissues,<br />

facilitate organ avoidance, and modify drug release profiles, minimizing toxicity. 91 The liposomal<br />

formulation of the popular anthracyline, doxorubicin, which is commonly used to treat metastatic<br />

breast and ovarian cancer, is reported to have diminished cardiotoxicity and enhanced therapeutic<br />

efficacy compared to the free form of the drug. 15–17 This increased efficacy is most likely due to the<br />

passive targeting of solid tumors through the enhanced permeability and retention effect inherent to<br />

tumor vasculature and aberrant tumor morphology. 17 The approximate diameter of the doxorubicin–liposomal<br />

product is reported to be 100 nm, near the size limits described in the FDA<br />

definition of a nanomedicine.<br />

Nanoparticlized, albumin-bound paclitaxel, Abraxane w , is one of the first FDA-approved<br />

(January, 2005) cancer drugs that was submitted to the FDA with specific claims regarding<br />

nano-related characteristics that improved this particular formulation over previous submissions.<br />

18–20 Initially identified as (ABI-007), this chemotherapeutic agent that is approved for<br />

the treatment of metastatic breast cancer eliminates the excipient Cremophor w <strong>by</strong> overcoming<br />

intrinsic insolubility with nanoparticlization of the active paclitaxel ingredient. Although widely<br />

used in many drug formulations, Cremophor has been associated with an adverse hypersensitivity<br />

reaction that limited the amount of paclitaxel that could be safely administered. 20 Unlike the<br />

previous three agents, the approximate diameter of Abraxane is 130 nm, slightly larger than the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 143<br />

requirements described in the National Nanotechnology Initiative (NNI) definition of nanotechnology<br />

(!100 nm), and thus originally was determined not to fit the size requirement of a<br />

nanomedicine. However, the fact that the exact particle size exceeds 100 nm may not necessarily<br />

preclude Abraxane and other similarly-sized or larger products from being<br />

considered nanotechnology.<br />

8.3 FDA DEFINITION OF NANOTECHNOLOGY<br />

As described on the NNI website (http://www.nano.gov/index.html), “Nanotechnology is the<br />

understanding and control of matter at dimensions of roughly 1–100 nm, where unique phenomena<br />

enable novel applications. Encompassing nanoscale science, engineering and technology, nanotechnology<br />

involves imaging, measuring, modeling, and manipulating matter at this length scale.”<br />

Similarly, in a recent publication <strong>by</strong> the European Science Foundation on Nanomedicine, 92 nanotechnology<br />

is defined as the following: “In case of these devices, nanoscale objects were defined as<br />

molecules or devices within the size range of one to hundreds of nanometers that are the active<br />

component or object, even within the frame-work of a larger micro-size device or at a macrointerface.”<br />

A discussion regarding the definition of nanotechnology is currently an active topic<br />

among various organizations. However, while a definition is always useful, the FDA has traditionally<br />

dealt with products on a case-<strong>by</strong>-case basis. This approach is likely to continue for<br />

nanotechnology products, regardless of the definition. However, we are actively involved in the<br />

various organizations and committees dealing with issues regarding terminology, such as the<br />

American Society for Testing and Materials (ASTM) International.<br />

The definition of nanotechnology at the FDA has been consistent with the definition of nanotechnology<br />

as reported <strong>by</strong> the NNI. As such, the FDA currently defines nanotechnology with the<br />

following definition:<br />

1. The existence of materials or products at the atomic, molecular, or macromolecular<br />

levels, where at least one dimension that affects the functional behavior of the drug/device<br />

product is in the length scale range of approximately 1–100 nm.<br />

2. The creation and use of structures, devices, and systems that have novel properties and<br />

functions because of their small size.<br />

3. The ability to control or manipulate the product on the atomic scale.<br />

8.4 FDA INITIATIVES IN NANOTECHNOLOGY<br />

The FDA is responsible for protecting the health of the public <strong>by</strong> assuring the safety, efficacy and<br />

security of human and veterinary drugs, biological products, medical devices, our nation’s food<br />

supply, cosmetics, and products that emit radiation. The FDA is also responsible for advancing the<br />

public’s health <strong>by</strong> helping to speed innovations that make medicines and foods more effective, safer<br />

and more affordable; and <strong>by</strong> helping the public get the accurate, science-based information they<br />

need to use medicines and foods to improve their health. Towards this end, the FDA needs to<br />

proactively pursue those innovations that would result in a significant enhancement of public<br />

health. Therefore, it falls directly within the mandate of its mission for the FDA to understand<br />

and help to speed innovations that promise to improve many products that will be regulated <strong>by</strong> the<br />

FDA, such as those resulting from nanotechnology.<br />

In response to a slowdown in innovating medical therapies submitted to the FDA for approval, the<br />

FDA released the Critical Path Initiative (http://www.fda.gov/oc/initiatives/criticalpath/whitepaper.<br />

html) in March 2004. The report describes the urgent need to modernize the medical product<br />

development process—the Critical Path—to make product development more predictable and less<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


144<br />

costly. The Critical Path Initiative at the FDA is a serious attempt to bring attention and focus to the<br />

need for targeted scientific efforts to modernize the techniques and methods used to evaluate the<br />

safety, efficacy, and quality of medical products as they move from product selection and design to<br />

mass manufacture. It is intended to integrate new science into the regulatory process. As such, a list of<br />

opportunities for the Critical Path Initiative was published <strong>by</strong> the FDA, and nanotechnology was<br />

identified as one of the elements under the Critical Path Initiative. Figure 8.1 illustrates the three<br />

dimensions of the Critical Path (safety, medical utility and industrialization).<br />

However, the Food and Drug Administration is only one of several government agencies that<br />

currently regulates and evaluates a wide range of products that utilize nanotechnology and/or<br />

contain nanomaterials, including foods, cosmetics, drugs, devices, and veterinary products. As a<br />

member of both the National Science and Technology Council (NSTC) Committee on Technology,<br />

and the Nanoscale Science, Engineering and Technology (NSET) Working <strong>Group</strong> on Nanomaterials<br />

Environmental and Health Implications (NEHI), the FDA coordinates with other government<br />

agencies, including the National Institute for Environmental Health Sciences (NIEHS), the<br />

National Toxicology Program (NTP), and the National Institute of Occupational Safety and<br />

Health (NIOSH) to develop novel strategies for the characterization of and safety evaluation<br />

standards for nanomaterials. The FDA also works closely with the National Cancer Institute<br />

(NCI) and the National Institute of Standards and Technology (NIST) in the Interagency Oncology<br />

Task Force, Nanotechnology subcommittee, to pool knowledge and resources to facilitate the<br />

development and approval of new cancer drugs and to address the use of nanotechnology in<br />

cancer therapies and diagnostics. Within the FDA, the Office of Science and Health Coordination<br />

manages the regular interaction and discussion of nanotechnology between the various agency<br />

centers and offices, and the nanotechnology working groups within each center, and address the<br />

concerns about the regulation of nanosized drugs and materials submitted to the individual centers.<br />

Dimensions<br />

Basic<br />

research<br />

Safety<br />

Medical<br />

utility<br />

Industrialization<br />

Prototype<br />

design or<br />

discovery<br />

Material selection<br />

structure<br />

activity<br />

relationship<br />

In vitro and<br />

computer<br />

model<br />

evaluation<br />

Physical<br />

design<br />

Preclinical<br />

development<br />

In vitro<br />

and animal<br />

testing<br />

In vitro<br />

and animal<br />

models<br />

Characterization<br />

small-scale<br />

production<br />

Nanotechnology for Cancer Therapy<br />

Clinical development<br />

Human<br />

and animal<br />

testing<br />

Human<br />

efficacy<br />

evaluation<br />

Manunfacturing<br />

scale-up<br />

refined<br />

specifications<br />

FDA filling/<br />

approval &<br />

launch<br />

preparation<br />

Safety<br />

follow<br />

up<br />

Mass<br />

production<br />

FIGURE 8.1 A highly generalized description of activities that must be successfully completed at different points<br />

and in different dimensions along the Critical Path. Many of these activities are highly complex—whole industries<br />

are devoted to supporting them. Not all the described activities are performed for every product, and many activities<br />

have been omitted for the sake of simplicity. (From http://www.fda.gov/oc/initiatives/criticalpath/whitepaper.<br />

html.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 145<br />

8.5 RESEARCH AT THE FDA ON NANOTECHNOLOGY<br />

Although the FDA is a regulatory agency, it is also engaged in a number of research activities.<br />

However, the FDA recognizes that science supported <strong>by</strong> research provides the foundation for sound<br />

regulation. As such, the FDA is interested in research projects that can address specific regulatory<br />

needs. As an example, FDA has a grants program in support of orphan products research and<br />

development. The FDA does not have a grants program to support other research in non-FDA<br />

laboratories. Nevertheless, in several of its Centers the FDA conducts research to understand the<br />

characteristics of nanomaterials and nanotechnology processes. Research interests include any<br />

areas related to the use of nanoproducts that the FDA needs to consider in the regulation of<br />

these products.<br />

Although many FDA Centers are engaged in some nanotechnology research activities, only<br />

some of the specific projects that are currently ongoing are listed below. The research projects,<br />

which happen to be conducted <strong>by</strong> the Center for Drug Evaluation and Research (CDER), include:<br />

1. Particle size determination in marketed sunscreens with TiO2 and ZnO nanoparticles.<br />

Sunscreens are considered drugs in the U.S., and although some undergo a premarket<br />

FDA review, others are marketed under over-the-counter (OTC) monographs that outline<br />

the active ingredients that the FDA has found to be safe and effective, as well as labeling<br />

that is truthful and not misleading. TiO2 and ZnO used in sunscreens are therefore<br />

manufactured according to OTC monographs, and these monographs do not specify<br />

particle size. As a result, a manufacturer can formulate a sunscreen using particles of<br />

TiO2 and ZnO that may or may not be in the nano size range. Therefore, to assess the<br />

particle size of TiO 2 and ZnO in currently marketed sunscreens, CDER has undertaken a<br />

research project, in collaboration with NIST.<br />

2. Another project focuses on manufacturing various nanoformulations in our laboratories<br />

and characterizing the physical and chemical properties of the nanoparticles in these<br />

formulations, including evaluating the effects of excipients on the measured parameters,<br />

the effects of preparation methodology on the measured parameters, and the impact (if<br />

any) that process and formulation variables may have on nanotechnology product<br />

characteristics and stability.<br />

3. Another project is studying the in vivo effects of selected nanoparticles, using validated<br />

animal models. The toxicity of the same nanoparticles will also be evaluated, using<br />

various in vitro systems.<br />

8.6 REGULATORY CONSIDERATIONS FOR NANOTECHNOLOGY PRODUCTS<br />

8.6.1 JURISDICTION OF NANOTECHNOLOGY PRODUCTS<br />

There has been much interest in the question of whether nanotechnology products will be<br />

considered drugs, devices, biologics or a combination of the three for regulatory purposes and<br />

assignment of work. This has been extensively discussed internally within the FDA and the current<br />

thinking is that many of these products will be considered combination products.<br />

Combination products (i.e., drug–device, drug–biologic, and device–biologic products) are<br />

increasingly incorporating cutting edge, novel technologies that hold great promise for advancing<br />

patient care. For example, innovative drug delivery devices have the potential to make treatments<br />

safer, more effective, or more convenient to patients. For example, drug-eluting cardiovascular<br />

stents have the potential to reduce the need for surgery <strong>by</strong> preventing the restenosis that sometimes<br />

occurs following stent implantation. Drugs and biologics can be used in combination to potentially<br />

enhance the safety and/or effectiveness of either product used alone. Biologics are being<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


146<br />

Nanotechnology for Cancer Therapy<br />

incorporated into novel orthopedic implants to help facilitate the regeneration of bone required to<br />

permanently stabilize the implants.<br />

Because combination products involve components that would normally be regulated under<br />

different types of regulatory authorities, and frequently <strong>by</strong> different FDA centers, they also raise<br />

challenging regulatory, policy, and review management issues. A number of criticisms have been<br />

raised regarding the FDA’s regulation of combination products. These include concerns about the<br />

consistency, predictability, and transparency of the assignment process; issues related to the<br />

management of the review process when two (or more) FDA centers have review responsibilities<br />

for a combination product; lack of clarity about the postmarket regulatory controls applicable to<br />

combination products; and lack of clarity regarding certain Agency policies, such as when applications<br />

to more than one agency center are needed.<br />

To address these concerns, FDA’s Office of Combination Products (OCP) was established on<br />

December 24, 2002, as required <strong>by</strong> the Medical Device User Fee and Modernization Act of 2002<br />

(MDUFMA). The law gives the office broad responsibilities covering the regulatory life cycle of<br />

drug–device, drug–biologic, and device–biologic combination products. However, the primary<br />

regulatory responsibilities for, and oversight of, specific combination products will remain in<br />

one of three product centers—the Center for Drug Evaluation and Research, the Center for<br />

Biologics Evaluation and Research, or the Center for Devices and Radiological Health—to<br />

which they are assigned.<br />

A combination product is assigned to an agency center or alternative organizational component<br />

that will have primary jurisdiction for its premarket review and regulation. Under section 503(g)(1)<br />

of the act, assignment to a center with primary jurisdiction, or a lead center, is based on a<br />

determination of the “primary mode of action” (PMOA) of the combination product. For<br />

example, if the PMOA of a device–biological combination product is attributable to the biological<br />

product, the agency component responsible for premarket review of that biological product would<br />

have primary jurisdiction for the combination product.<br />

A final rule defining the PMOA of a combination product was published in the August 25, 2005<br />

edition of the Federal Register, and is available at http://www.fda.gov/oc/combination/. The final<br />

rule defines PMOA as “the single mode of action of a combination product that provides the most<br />

important therapeutic action of the combination product.”<br />

8.6.2 EXISTING FDA REGULATIONS THAT APPLY TO NANOMATERIAL-CONTAINING PRODUCTS<br />

The FDA has specific guidance/requirements in place for most products. These guidance documents<br />

can be accessed on the FDA’s website (http://www.fda.gov/oc/industry/default.htm). These<br />

requirements apply to all products, whether they contain nanomaterials or not. Existing requirements<br />

are therefore considered to be adequate at this time for most nanotechnology<br />

pharmaceutical products.<br />

In the following section, there will be a brief description of the current preclinical requirements<br />

for the submission of most drugs, to highlight why we believe that our current regulations are<br />

adequate for the evaluation of the types of nanotechnology products that have been reported as<br />

being “close” to submission to the agency as an investigational new drug (IND).<br />

Within CDER, the preclinical requirements for approval to market pharmaceutical products<br />

include both short-term and long-term toxicity testing in rodent and non-rodent species. Specifically,<br />

the types of studies conducted <strong>by</strong> pharmaceutical manufacturers prior to New Drug<br />

Application (NDA) submission include pharmacology (mechanism of action), safety pharmacology<br />

(functional studies in various organ systems, most notably the cardiovascular system), absorption,<br />

distribution, metabolism, and excretion (ADME), genotoxicity (potential to cause mutations in both<br />

in vivo and in vitro assays), developmental toxicity (to assess effects on reproduction, fertility and<br />

lactation), irritation studies (to assess local irritation effects), immunotoxicology (to assess effects<br />

on the components of the immune system), carcinogenicity (to assess the capacity of drugs to<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 147<br />

induce tumors in animal models) and other possible studies (specific studies for a product being<br />

developed). The current battery of tests, as listed above, is therefore considered to be adequate<br />

because of the following reasons: the studies use high dose multiples of the drug (usually three<br />

doses, one that results in low, one that results in medium and one that results in high toxicity), at<br />

least two animal species are used (usually one rodent and one non-rodent), extensive histopathology<br />

is conducted on most organs (where most organs are removed, fixed, stained and viewed under the<br />

microscope to assess if there are structural changes that may have resulted from damage to the<br />

organs), functional tests are conducted to assess if there are effects on specific organ systems (such<br />

as the heart, brain, respiratory system, reproductive system, immune system, etc.) and drug treatments<br />

in animals can be for extended periods of time (up to two years for carcinogenicity studies).<br />

Additional studies might be requested based on drug-specific considerations and on a case<strong>by</strong>-case<br />

basis. Therefore, as for other drugs, nanotechnology products will be handled on a case<strong>by</strong>-case<br />

basis, depending on the characteristics of the particular product being developed. As<br />

research provides additional understanding with regards to the toxicity of nanomaterials, the<br />

FDA may require additional toxicological tests to ensure the safety of the products it regulates.<br />

8.7 SCIENTIFIC CONSIDERATIONS FOR THE DEVELOPMENT<br />

OF NANOTECHNOLOGY PRODUCTS<br />

Ensuring the safety of nanomaterials in drug applications is one of the most important concerns for<br />

regulators and drug developers. However, as important as safety is, it is also important to<br />

adequately characterize the nanomaterials used in drug products. Proper characterization and<br />

manufacturing procedures will ensure that a consistent product is being used, both during preclinical<br />

and clinical development, and after approval. Some of the questions that have been raised<br />

internally regarding the assurance of safety of nanomaterial-containing products, as well as questions<br />

regarding characterization, are listed below. It is not within the scope of this chapter to address<br />

questions regarding manufacturing and scale-up; however, it should be noted that this is an issue<br />

that needs serious consideration, because unique characteristics of nanotechnology products are<br />

likely to require unique manufacturing procedures.<br />

8.7.1 CHARACTERIZATION OF NANOPARTICLES<br />

It has become clear to most of those involved in research or development of potential nanotechnology<br />

applications that it is only possible to assess the safety of nanomaterials when the material<br />

under study is adequately characterized. If two or more laboratories are working on what they think<br />

is the same material, then in order for the data from one laboratory to be compared with data from<br />

another laboratory, there needs to be some confirmation that in fact the nanomaterials being studied<br />

are actually identical. It has been mentioned <strong>by</strong> many that for nanomaterials, small differences in<br />

product characteristics (such as size, surface charge, hydrophobicity, or a number of other attributes)<br />

may impact product behavior and thus result in vastly different safety profiles. Published data<br />

show that slight modifications of a nanoparticle have resulted in different toxicity profiles. For<br />

example, several in vitro studies with Starburst w polyamidoamine (PAMAM) dendrimers and<br />

Caco-2 cells have shown decreased cytotoxicity of cationic dendrimers upon addition of lauroyl<br />

or polyethylene glycol (PEG) to the surface of the macromolecules. 93–94 Similar findings of diminished<br />

toxicity were observed in J774 macrophages upon conjugation of polysaccharides to polymer<br />

nanoparticles and in fibroblasts and CHO cells with the addition of PEG–silica to the surface of<br />

CdSe and CdSe/ZnS nanocrystals. 95–96<br />

Although much less toxicity data is available in vivo, it is evident that surface chemistry is of<br />

similar importance to the toxico- and pharmacokinetics, toxico- and pharmacodynamics, and<br />

overall stability of the nanomaterials in vivo. 97–99 When injected into rodents, several generations<br />

of dendrimers with different surface configurations displayed contrasting organ distribution<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


148<br />

Nanotechnology for Cancer Therapy<br />

patterns and plasma clearance profiles in vivo, typically depositing within the liver, kidney, and<br />

pancreas. 98–100 PEG-conjugation to the dendrimer particle surface significantly increased the blood<br />

half-life and diminished the liver accumulation of dendrimer particles. 99 Similarly, quantum dots<br />

coated with a complex polymer with carbon alkyl side chains demonstrated greater in vivo stability<br />

and less genotoxicity than those coated with a simpler polymer or lipid coating. 101 Based on the<br />

findings of these studies and the collection of pulmonary inhalation data with environmental<br />

ultrafines and other in vitro cytotoxicity results obtained with different nanomaterials (e.g.,<br />

quantum dots, fullerenes, and nanoshells, etc.), it appears that a thorough characterization of the<br />

nanomaterial is essential to the interpretation and understanding of toxicity data collected from<br />

in vitro and in vivo studies.<br />

To summarize some of the ongoing discussions within CDER’s nanotechnology working<br />

group, a list of questions has been put together regarding the characterization of nanomaterialcontaining<br />

products. At this time, it appears that there are no adequate answers to these questions.<br />

However, it may be that in the future, to satisfy regulatory requirements for nanomaterialcontaining<br />

products, the questions listed below will need to be addressed <strong>by</strong> product sponsors.<br />

1. What are the forms in which particles are presented to the organism, the cells and the<br />

organelles? Are these particles soluble or insoluble, are the nanomaterials organic or<br />

inorganic molecules, are the nanomaterials described as nanoemulsions, nanocrystal<br />

colloid dispersions, liposomes, nanoparticles that are combination products (drug–<br />

device, drug–biologic, drug–device–biologic), or something else? Does the product<br />

fall within the currently established definition of nanotechnology?<br />

2. What are the tools that can be used to characterize the properties of the nanoparticles?<br />

Can standard tools that are used of other types of products (for example, spectroscopic<br />

tools) be used to characterize nanomaterial-containing products, or do these products<br />

require specialized tools that are not widely available for product characterization<br />

(atomic force microscopy, tunneling electron microscopy, or other specialized modalities)?<br />

3. Are there validated assays to detect and quantify nanoparticles in the drug product and in<br />

tissues?<br />

4. How can long- and short-term stability of nanomaterials be assessed? Specifically, how<br />

can the stability of nanomaterials be assessed in various environments (buffer, blood,<br />

plasma, and tissues)?<br />

5. What are the critical physical and chemical properties of nanomaterials in a drug product<br />

and how do residual solvents, processing variables, impurities and excipients impact<br />

these properties?<br />

6. How do physical characteristics impact product quality and performance? Within what<br />

range can these physical properties be modified without significantly affecting product<br />

quality and performance?<br />

7. What are the critical steps in the scale-up and manufacturing process for nanotechnology<br />

products? How could manufacturing procedures for nanomaterials differ from those for<br />

other types of drugs?<br />

8. How are issues concerning the characterization and manufacturing procedures for nanotechnology<br />

products assessed, when considering the development of “personalized<br />

therapies”? In this context, personalized therapies refer to those multifunctional products<br />

that will be “custom made” for specific patients, depending on issues such as the<br />

expression of a certain receptor on a tumor or the response of the tumor to a particular<br />

chemotherapeutic agent. For these types of therapies, what should be the level of<br />

characterization of the nanomaterials? Does the preclinical testing need to be repeated<br />

for each modification of the multifunctional molecule, or can there be limited testing to<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 149<br />

cover certain aspects of the product behavior, such as the ADME, the toxicology (a<br />

bridging study for the acute and/or repeat dose toxicology studies)? What aspects of<br />

the chemistry, manufacturing and controls (CMC) studies need to be repeated and what<br />

should be the extent of physical characterization?<br />

8.7.2 SAFETY CONSIDERATIONS<br />

Other than characterization, the other major consideration has to do with safety. However, it is only<br />

within the context of adequately characterized products that safety studies provide value. Therefore,<br />

although safety is instinctively the primary concern, adequate characterization of<br />

nanomaterials has to precede any studies on safety.<br />

As particle size decreases, there may be size-specific effects. These effects may impact safety if<br />

the materials are more reactive. It has been mentioned in numerous discussions that, as particle size<br />

decreases, reactivity is expected to rise. The increased reactivity combined with the decreased size<br />

has led us to pose many questions. Some of these questions are listed below:<br />

1. Will nanoparticles gain access to tissues and cells that normally would be <strong>by</strong>passed <strong>by</strong><br />

larger particles?<br />

2. If nanoparticles enter cells, what effects do they have on cellular and tissue functions?<br />

Are the effects transient or permanent? Might there be different effects in different cell<br />

types, depending on specific tissue targets or receptors?<br />

3. Once nanoparticles enter tissues, how long do they remain there and where do they<br />

concentrate?<br />

4. What are the mechanisms <strong>by</strong> which nanoparticles are cleared from tissues and blood and<br />

how can their clearance be measured?<br />

5. What are some route-specific issues for nanomaterials, and how are these issues different<br />

than for other types of materials that are not in the nanoscale?<br />

6. With regards to ADME, specific questions include:<br />

a. What are the differences in the ADME profile for nanoparticles versus larger particles<br />

of the same drug?<br />

b. Are current methods used for measuring drug levels in blood and tissues adequate for<br />

assessing levels of nanoparticles (appropriateness of method, limits of detection)?<br />

c. How accurate are mass balance studies, especially if the levels of drug administered<br />

are very low; i.e., can 100% of the amount of the administered drug be accounted for?<br />

d. How is clearance of targeted nanoparticles accurately assessed?<br />

e. If nanoparticles concentrate in a particular tissue, how can clearance be accurately<br />

determined?<br />

f. Can nanoparticles be successfully labeled for ADME studies?<br />

8.8 CONCLUSIONS<br />

In this chapter, we have tried to highlight some of the steps being taken <strong>by</strong> the FDA to meet the<br />

challenges of nanotechnology products. The FDA has engaged in an open discussion with numerous<br />

groups and organizations, has participated in and sponsored workshops, has established internal<br />

working groups, has outlined a path for the review of combination nanotechnology products and has<br />

initiated research projects in nanotechnology. All these efforts have been geared towards a better<br />

understanding of the science and towards assessing the adequacy of existing regulations for the<br />

types of products that are expected to be regulated <strong>by</strong> the FDA.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


150<br />

At this point, and based on internal discussions, the FDA feels that its current regulations are<br />

adequate to address the types of nanotechnology products that are being developed as drugs or drug<br />

delivery systems. The reasons for this have been outlined in this chapter. Although there have been<br />

requests <strong>by</strong> investigators and developers of nanotechnology products for the FDA to issue nanotechnology-specific<br />

guidances, no new guidance documents applicable specifically to<br />

nanotechnology products will be issued at this time. This is because it is felt that all existing<br />

Guidance documents apply to these products. All products being developed, whether nanotechnology-based<br />

or not, are unique and therefore likely to have unique issues requiring consideration.<br />

As such, and as for all other products reviewed <strong>by</strong> the FDA, nanotechnology products will be<br />

reviewed on a case-<strong>by</strong>-case basis.<br />

As shown in Figure 8.2, the drug development process is multiphasic and along the path, there<br />

are numerous opportunities for meetings between the sponsor and the FDA. The first meeting is the<br />

pre-IND meeting, and it is highly recommended that sponsors of nanotechnology products take the<br />

opportunity to meet with the FDA at this early stage. An early meeting with the FDA can avoid<br />

unnecessary or unfocussed studies that could result in a waste of sponsor resources. During early<br />

meetings, sponsors can present, with minimal data, their preclinical plans to the FDA and obtain<br />

guidance on how to focus the studies submitted in the IND package. This type of interaction<br />

between the FDA and the sponsor will be valuable to both parties and will pave the way for the<br />

most efficient drug development plan.<br />

Basic<br />

research<br />

Industry - FDA<br />

interactions<br />

during<br />

development<br />

Prototype<br />

design or<br />

discovery<br />

Preclinical<br />

development<br />

Pre-IND meeting<br />

Initial<br />

IND<br />

submissions<br />

Ongoing<br />

submission<br />

End of<br />

phase<br />

2a meeting<br />

Nanotechnology for Cancer Therapy<br />

Clinical development<br />

Phase 1 Phase 2 Phase 3<br />

End of phase<br />

2 meeting<br />

Pre-BLA or NDA<br />

meeting<br />

FDA filing/<br />

approval &<br />

launch<br />

preparation<br />

Safety update<br />

Market<br />

application<br />

submission<br />

IND review phase Application<br />

review<br />

phase<br />

FIGURE 8.2 This figure depicts the extensive industry–FDA interactions that occur during product development,<br />

using the drug development process as a specific example.* <strong>Developers</strong> often meet with the agency<br />

before submitting an investigational new drug (IND) application to discuss early development plans. An IND<br />

must be filed and cleared <strong>by</strong> the FDA before human testing can commence in the United States. During the<br />

clinical phase, there are ongoing submissions of new protocols and results of testing. <strong>Developers</strong> often request<br />

additional meetings to get FDA agreement on the methods proposed for evaluation of safety or efficacy, and<br />

also on manufacturing issues. *Note: Clinical drug development is conventionally divided into three phases.<br />

This is not the case for medical device development. (From www.fda.gov/oc/initiatives/criticalpath/whitepaper.html.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 151<br />

8.9 TRADEMARK LISTING<br />

1. Magnevist w (Schering Aktiengesellschaft, Berlin, Germany)<br />

2. Feridex w (Advanced Magnetics, Inc., Cambridge, Massachusetts, U.S.A.)<br />

3. Rapamune w (Wyeth, Madison, New Jersey, U.S.A.)<br />

4. Emend w (Merck & CO., Inc., Whitehouse Station, New Jersey, U.S.A.)<br />

5. Doxil w (Ortho Biotech Products, L.P., Bridgewater, New Jersey, U.S.A.)<br />

6. Abraxane w (American Pharmaceutical Partners, Inc., Schaumburg, Illinois, U.S.A.)<br />

7. SilvaGard w (AcryMed, Beaverton, Oregon, U.S.A.)<br />

8. AmBisome w (Gilead Sciences, Inc., Foster City, California, U.S.A.)<br />

9. Leunesse w (Nanotherapeutics, Inc., Alachua, Florida, U.S.A.)<br />

10. VivaGel w (Starpharma Holdings, Ltd., Melbourne, Australia)<br />

11. Combidex w (Advanced Magnetics, Inc., Cambridge, Massachusetts, U.S.A.)<br />

12. Cremophor w (BASF Aktiengesellschaft, CO., Ludwigshafen, Germany)<br />

13. Starburst w (Dendritic Nanotechnologies, Inc., Mount Pleasant, Michigan, U.S.A.)<br />

REFERENCES<br />

1. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Nanomedicine: Current status and future<br />

prospects, FASEB Journal, 19 (311), 330, 2005.<br />

2. Brigger, I., Dubernet, C., and Couvreur, P., Nanoparticles in cancer therapy and diagnosis, Advanced<br />

Drug Delivery Review, 54 (5), 631–651, 2002.<br />

3. Kieffer, S. A., Gadopentetate dimeglumine: Observations on the clinical research process,<br />

Radiology, 174 (1), 7–8, 1990.<br />

4. Zhang, H. L., Ersoy, H., and Prince, M. R., Effects of gadopentetate dimeglumine and gadodiamide<br />

on serum calcium, magnesium, and creatinine measurements, Journal of Magnetic Resonance<br />

Imaging, 23, 383–387, <strong>2006</strong>.<br />

5. de Lussanet, Q. G., Backes, W. H., Griffioen, A. W., van Engleshoven, J. M., and Beets-Tan, R. G.,<br />

Gadopentenate dimeglumine versus ultra-small superparamagnetic iron oxide for dynamic contrastenhanced<br />

MR imaging of tumor angiogenesis in human colon carcinoma in mice, Radiology, 229,<br />

429–438, 2003.<br />

6. Frank, J. A., Zywicke, H., Jordan, E. K., Mitchell, J., Lewis, B. K., Miller, B., Bryant, L. H. Jr., and<br />

Bulte, J. W., Magnetic intracellular labeling of mammalian cells <strong>by</strong> combining (FDA approved)<br />

superparamagnetic iron oxide MR contrast agents and commonly used transfection agents, Academy<br />

of Radiology, 9 (Suppl. 2), S484–S487, 2002.<br />

7. Shapiro, E. M., Skrtic, S., Sharer, K., Hill, J. M., Dunbar, C. E., and Koretsky, A. P., MRI detection of<br />

single particles for cellular imaging, Proceedings of the National Academy of Sciences of the United<br />

States of America, 101 (30), 10901–10906, 2004.<br />

8. Daldrup-Link, H. E., Rudelius, M., Oostendorp, R. A., Settles, M., Piontek, G., Metz, S.,<br />

Rosenbrock, H., Keller, U., Heinzmann, U., Rummeny, E. J., Schlegel, J., and Link, T. M., Targeting<br />

of hematopoietic progenitor cells with MR contrast agents, Radiology, 228 (3), 760–767, 2003.<br />

9. Kostura, L., Kraitchman, D. L., Macka, A. M., Pittenger, M. F., and Bulte, J. W., Feridex labeling of<br />

mesenchymal stem cells inhibits chondrogenesis but not adipogenesis or osteogenesis, NMR in<br />

Biomedicine, 17 (7), 513–517, 2004.<br />

10. LaVan, D. A., Lynn, D. M., and Langer, R., Moving smaller in drug discovery and delivery, National<br />

Review of Drug Discovery, 1, 77–84, 2002.<br />

11. Miller, J. L., Sirolimus approved with renal transplant indication, American Journal of Health-<br />

System Pharmacy, 56 (21), 2177–2178, 1999.<br />

12. Aspeslet, L. J. and Yatscoff, R. W., Requirements for therapeutic drug monitoring of sirolimus, an<br />

immunosuppressive agents used in renal transplant, Clinical Therapeutics, 22 (Suppl. B), B86–B92,<br />

2000.<br />

13. Patel, L. and Lindley, C., Aprepitant-a novel NK1-receptor antagonist, Expert Opinion on Pharmacotherapy,<br />

4 (12), 2279–2296, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


152<br />

Nanotechnology for Cancer Therapy<br />

14. Wu, Y., Loper, A., Landis, E., Hettrick, L., Novak, L., Lynn, K., Chen, C., Thompson, K., Higgins,<br />

R., Batra, U., Shelukar, S., Kwei, G., and Storey, D., The role of biopharmaceutics in the development<br />

of a clinical nanoparticle formulation of MK-0869: A beagle dog model predicts improved<br />

bioavailability and diminished food effect on absorption in human, International Journal of Pharmaceutics,<br />

285 (1–2), 135–146, 2004.<br />

15. Markman, M., Kennedy, A., Webster, K., Peterson, G., Kulp, B., and Belinson, J., Phase 2 trial of<br />

liposomal doxorubicin (40 mg/m(2)) in platinum/paclitaxel refractory ovarian fallopian tube cancers<br />

and primary carcinoma of the peritoneum, Gynecologic Oncology, 78, 369–372, 2000.<br />

16. Frykman, G., Williams, G., and Pazdur, R., Conflicting phase II efficacy date for Doxil, Journal of<br />

Clinical Oncology, 19 (2), 596–597, 2001.<br />

17. Laginha, K. M., Verwoert, S., Charrois, G. J. R., and Allen, T. M., Determination of doxorubicin<br />

levels in whole tumor and tumor nuclei in murine breast cancer tumors, Clinical Cancer Research,11<br />

(19), 6944–6949, 2005.<br />

18. Harries, M., Ellis, P., and Harper, P., Nanoparticle albumin-bound paclitaxel for metastatic breast<br />

cancer, Journal of Clinical Oncology, 23 (31), 7768–7771, 2005.<br />

19. Nyman, D. W., Campbell, K. J., Hersh, E., Long, K., Richardson, K., Trieu, V., Desai, N., Hawkins,<br />

M. J., and Von Hoff, D. D., Phase I and pharmacokinetics trial of ABI-007, a novel nanoparticle<br />

formulation of paclitaxel in patients with advanced nonhematologic malignancies, Journal of<br />

Clinical Oncology, 23 (31), 7785–7793, 2005.<br />

20. Sparreboom, A., Scripture, C. D., Trieu, V., Williams, P. J., De, T., Yang, A., Beals, B., Figg, W. D.,<br />

Hawkins, M., and Desai, N., Comparative preclinical and clinical pharmacokinetics of a cremaphorfree<br />

nanoparticle albumin-bound paclitaxel (ABI-007) and Paclitaxel formulated in cremaphor<br />

(Taxol), Clinical Cancer Research, 11 (11), 4136–4143, 2005.<br />

21. Chaw, K. C., Manimaran, M., and Tay, F. E., Role of silver ions in destabilization of intermolecular<br />

adhesion forces measured <strong>by</strong> atomic force microscopy in staphylococcus epidermis biofilms, Antimicrobial<br />

Agents and Chemotherapy, 49 (12), 4853–4859, 2005.<br />

22. Warriner, R. and Burrell, R., Infection and the chronic wound: A focus on silver, Advanced Skin<br />

Wound Care, 18 (Suppl. 1), 2–12, 2005.<br />

23. Baker, C., Pradhan, A., Pakstis, L., Pochan, D. J., and Shah, S. I., Synthesis and antibaceterial<br />

properties of silver nanoparticles, Journal of Nanoscience and Nanotechnology, 5 (2), 244–249,<br />

2005.<br />

24. Takemoto, K., Yamamoto, Y., Ueda, Y., Sumita, Y., Yoshida, K., and Niki, Y., Comparative study<br />

on the efficacy of Ambisome and Fungizone in a mouse model of pulmonary aspergillosis, Journal of<br />

Antimicrobial Chemotherapy, 57 (4), 724–731, <strong>2006</strong>.<br />

25. Cetin, H., Yalaz, M., Akisu, M., Hilmioglu, S., Metin, D., and Kultursay, N., The efficacy of two<br />

different lipid-based amphotericin B in neonatal Candida septicemia, Pediatrics International, 47<br />

(6), 676–680, 2005.<br />

26. Barratt, G. and Legrand, P., Comparison of the efficacy and pharmacology of formulations of<br />

amphotericin B used in treatment of leishmaniasis, Current Opinion in Infectious Diseases, 18<br />

(6), 527–530, 2005.<br />

27. The Healthcare Sales and Marketing Network. NanoBio w corporation announces FDA approval for<br />

phase II clinical trials for herpes labialis treatment. http://salesandmarketingnetwork.com/news_release.php?IDZ14361<br />

(accessed February 24, 2004).<br />

28. NewsRX. FDA approves phase II clinical trials for herpes labialis treatment. http://www.newsrx.<br />

com/article.php?articleIDZ87034 (accessed February 24, 2004).<br />

29. BioSpace. NanoBio announces FDA approval for phase II clinical trials for herpes labialis treatment.<br />

http://www.biospace.com/news_story.aspx?StoryIDZ15257520&fullZ1 (accessed February 24,<br />

2004).<br />

30. Nanotsunami. NanoBio w completes successful phase 2 herpes trial, prepares for phase 3 trial. http://<br />

www.voyle.net/Nano%20Medicine%202005/Medicine%202005-0087.htm (accessed July 22,<br />

2005).<br />

31. Abner, S. R., Guenthner, P. C., Guarner, J., Hancock, K. A., Cummins, J. E. Jr., Fink, A., Gilmore,<br />

G. T., Staley, C., Ward, A., Ali, O., Binderow, S., Cohen, S., Grohskopf, L. A., Paxton, L., Hart,<br />

C. E., and Dezzutti, C. S., A human colorectal explant culture to evaluate topical microbicides for the<br />

prevention of HIV infection, Journal of Infectious Diseases, 192 (9), 1545–1556, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 153<br />

32. Gong, E., Matthews, B., McCarthy, T., Chu, J., Holan, G., Raff, J., and Sacks, S., Evaluation of<br />

dendrimer SPL7013, a lead microbicide candidate against herpes simplex virus, Antiviral Research,<br />

68 (3), 139–146, 2005.<br />

33. McCarthy, T. D., Karellas, P., Henderson, S. A., Giannis, M., O’Keefe, D. G., Heery, G., Paull,<br />

J. R. A., Matthews, B. R., and Holan, G., Dendrimers as drugs, discovery and preclinical and clinical<br />

development of dendrimer-based microbicides for HIV and STI prevention, Molecular Pharmacology,<br />

2 (4), 312–318, 2005.<br />

34. Jiang, Y. H., Emau, P., Cairns, J. S., Flanary, L., Morton, W. R., McCarthy, T. D., and Tsai, C. C.,<br />

SPL7013 gel as topical microbicide for prevention of vaginal transmission of SHIV89.6P in macaques,<br />

AIDS Research and Human Retroviruses, 21 (3), 207–213, 2005.<br />

35. Ito, I., Ji, L., Tanaka, F., Saito, Y., Gopalan, B., Branch, C. D., Xu, K., Atkinson, E. N., Bekele, B. N.,<br />

Stephens, L. C., Minna, J. D., Roth, J. A., and Ramesh, R., Liposomal vector mediated delivery of the<br />

3p FUS1 gene demonstrates potent antitumor activity against human lung cancer in vivo, Cancer<br />

Gene Therapy, 11 (11), 733–739, 2004.<br />

36. Smalltimes. Study: Nano therapy kills tumors, extends life. http://www.smalltimes.com/document_display.cfm?section_idZ53&document_idZ8429<br />

(accessed November 11, 2004).<br />

37. Nanotechnology News. Introgen’s nanoparticle tumor treatment increases survival rate <strong>by</strong> 70%.<br />

http://www.azonano.com/news.asp?newsIDZ406 (accessed November 12, 2004).<br />

38. Nanotsunami. Introgen’s nanoparticle cancer therapy INGN 401 demonstrates promise in the treatment<br />

of lung cancer. http://www.voyle.net/Nano%20Medicine%202005/Medicine%202005-0052.<br />

htm (accessed May 17, 2005).<br />

39. Bourrinet, P., Bengele, H. H., Bonnemain, B., Dencausse, A., Idee, J. M., Jacobs, P. M., and Lewis,<br />

J. M., Preclinical safety and pharmacokinetic profile of ferumoxtran-10, an ultrasmall superparamagnetic<br />

iron oxide magnetic resonance contrast agent, Investigative Radiology, 41 (3), 313–324,<br />

<strong>2006</strong>.<br />

40. Saksena, M. A., Saokar, A., and Harisinghani, M. G., Lymphotropic nanoparticle enhanced MR<br />

imaging (LNMRI) technique for lymph node imaging, European Journal of Radiology, 58 (3),<br />

367–374, <strong>2006</strong>.<br />

41. Harisinghani, M. G., Saksena, M. A., Hahn, P. F., King, B., Kim, J., Torabi, M. T., and Weissleder,<br />

R., Ferumoxtran-10-enhanced MR lymphangiography: Does contrast-enhanced imaging alone<br />

suffice for accurate lymph node characterization?, American Journal of Roentgenology, 186 (1),<br />

144–148, <strong>2006</strong>.<br />

42. Metz, S., Lohr, S., Settles, M., Beer, A., Woertler, K., Rummeny, E. J., and Daldrup-Link, H. E.,<br />

Ferumoxtran-10-enhanced MR imaging of the bone marrow before and after conditioning therapy in<br />

patients with non-Hodgkin lymphomas, European Radiology, 16 (3), 598–607, <strong>2006</strong>.<br />

43. Schluep, T., Hwang, J., Cheng, J., Heidel, J. D., Barlett, D. W., Hollister, B., and Davis, M. E.,<br />

Preclinical efficacy of the camptothecin-polymer conjugate IT-101 in multiple cancer models,<br />

Clinical Cancer Research, 12 (5), 1606–1614, <strong>2006</strong>.<br />

44. Schluep, T., Cheng, J., Khin, K. T., and Davis, M. E., Pharmacokinetics and biodistribution of the<br />

camptothecin–polymer conjugate IT-101 in rats and tumor-bearing mice, Cancer Chemotherapy and<br />

Pharmacology, 57, 654–662, <strong>2006</strong>.<br />

45. Newkome, G. R., Yao, Z., Baker, G. R., and Gupta, V. K., Cascade molecules: A new approach to<br />

micelles. A (27)-Arborol, Journal of Organic Chemistry, 50, 2003–2004, 1985.<br />

46. Tomalia, D. A., Baker, H., Dewald, J. R., Hall, M., Kallos, G., Maris, S., Roeck, J., Ryder, J., and<br />

Smith, P. A., New class of polymers: Starburst–dendritic macromolecules, Polymer Journal (Tokyo),<br />

17, 117–123, 1985.<br />

47. Wiener, E. C., Brechbiel, M. W., Brothers, H., Magin, R. L., Gansow, O. A., Tomalia, D. A., and<br />

Lauterbur, P. C., Dendrimer-based metal chelates: A new class of magnetic resonance imaging<br />

contrast agents, Magnetic Resonance in Medicine, 31 (1), 1–8, 1994.<br />

48. Grinstaff, M. W., Biodendrimers: New polymeric biomaterials for tissue engineering, Chemistry, 8,<br />

2839–2846, 2002.<br />

49. Bourne, N., Stanberry, L. R., Kern, E. R., Holan, G., Matthews, B., and Berstein, D. I., Dendrimers, a<br />

new class of candidate topic microbicides with activity against herpes simplex virus infestation,<br />

Antimicrobial Agents in Chemotherapy, 44, 2471–2474, 2000.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


154<br />

Nanotechnology for Cancer Therapy<br />

50. Zanini, D. and Roy, R., Practical synthesis of PAMAM athiosialodendrimers for probing multivalent<br />

carbohydrate–lectin binding properties, Journal of Organic Chemistry, 63, 3486–3491, 1998.<br />

51. Zhuo, R. X., Du, B., and Lu, Z. R., In vitro release of 5-fluoroacil with cyclic core dendritic polymer,<br />

Journal of Controlled Release, 57, 249–257, 1999.<br />

52. Liu, M., Kono, K., and Frechet, J. M. J., Water-soluble dendritic unimolecular micelles: Their<br />

potential as drug delivery agents, Journal of Controlled Release, 65, 121–131, 2000.<br />

53. Kobayashi, H. and Brechbiel, M. W., Dendrimer-based nanosized MRI contrast agents, Current<br />

Pharmaeutical Technology, 5, 539–549, 2004.<br />

54. Konda, S. D., Wang, S., Brechbiel, M., and Wiener, E. C., Biodistribution of a 153 Gd-folate<br />

dendrimer generationZ4, in mice with folate-receptor positive and negative ovarian tumor xenografts,<br />

Investigative Radiology, 37, 199–204, 2002.<br />

55. Hirsch, L. R., Stafford, R. J., Bankson, J. A., Sershen, S. R., Rivers, B., Price, R. E., Hazle, J. D.,<br />

Halas, N. J., and West, J. L., Nanoshell-mediated nar-infared thermal therapy of tumors under<br />

magnetic resonance guidance, Proceedings of the National Academy of Sciences of the United<br />

States of America, 100, 13549–13554, 2003.<br />

56. Loo, C., Lin, A., Hirsch, L., Lee, M. H., Barton, J., Halas, N., West, J., and Drezek, R., Nanoshellenabled<br />

photonics-based imaging and therapy of cancer, Technology Cancer Research and Treatment,<br />

3, 33–40, 2004.<br />

57. Loo, C., Lowery, A., Halas, N., West, J., and Drezek, R., Immunotargeted nanoshells for integrated<br />

cancer imaging and therapy, Nano Letters, 5, 709–711, 2005.<br />

58. O’Neal, D. P., Hirsch, L. R., Halas, N., Payne, J. D., and West, J. L., Photo-thermal tumor ablation in<br />

mice using near-infrared-absorbing nanoparticles, Cancer Letters, 209 (2), 171–176, 2004.<br />

59. de Jong, W. H., Roszek, B., and Geertsman, R. E., Nanotechnology in medical applications: Possible<br />

risks for human health. RIVM Report 265001002, Bilthoven, 2005.<br />

60. Haddon, R. C., Hebard, A. F., Rosseinsky, M. J., Murphy, D. W., Duclos, S. J., Lyons, K. B., Miller,<br />

B., Rosamilia, J. M., Fleming, R. M., Kortan, A. R., Glarum, S. H., Makjiha, A. V., Muller, A. J.,<br />

Eick, R. H., Zahurak, S. M., Tycko, R., Dabbagh, G., and Thiel, F. A., Conducting films of C60 and<br />

C70 <strong>by</strong> alkali-metal doping, Nature, 350, 320–322, 1991.<br />

61. Bosi, S., Da Ros, T., Spalluto, G., and Prato, M., Fullerene derivatives: An attractive tool for<br />

biological applications, European Journal of Medical Chemistry, 38, 913–923, 2003.<br />

62. Friedman, S. H., Decamp, D. L., Sijbesma, R. P., Srdanov, G., Wudl, F., and Kenyon, G. L.,<br />

Inhibition of the HIV-1 protease <strong>by</strong> fullerene derivatives: Model building studies and experimental<br />

verification, Journal of the American Chemical Society, 115, 6506–6509, 1993.<br />

63. Schiniazi, R. F., Sijbesma, R., Srdanov, G., Hill, C. L., and Wudl, F., Synthesis and virucidal activity<br />

of a water soluble, configurationally stable, derivatized C60 fullerene, Antimicrobial Agents and<br />

Chemotherapy, 37, 1707–1710, 1993.<br />

64. Chueh, S. C., Lai, K. M., Lee, M. S., Chiang, L. Y., Ho, T., and Chen, S. C., Decrease of free radical<br />

level in organ perfusate <strong>by</strong> a novel water soluble carbon-sixty hexa(sulfobutyl)fullerenes, Transplant<br />

Proceedings, 31, 1976–1977, 1999.<br />

65. Lin, A. M., Chyi, B. Y., Wang, S. D., Yu, H., Kanakamma, P. P., Luh, T. Y., Chou, C. K., and Ho,<br />

L. T., Carboxyfullerene prevents iron-induced oxidative stress in rat brain, Journal of Neurochemistry,<br />

72, 1634–1640, 1999.<br />

66. Cagle, D. W., Kennel, S. J., Mirzadeh, S., Alford, J. M., and Wilson, L. J., In vivo studies of<br />

fullerene-based materials using endohedral metallofullerene radiotracers, Proceedings of the<br />

National Academy of Sciences of the United States of America, 96, 5182–5187, 1999.<br />

67. Gharbi, N., Pressac, M., Hadchouel, M., Szwarc, H., Wilson, S. R., and Moussa, F., (60)Fullerene is a<br />

powerful antioxidant in vivo with no acute or subacute toxicity, Nano Letters, 5 (12), 2578–2585,<br />

2005.<br />

68. Bruchez, M. Jr., Moronne, M., Gin, P., Weiss, S., and Alivisatos, A. P., Semiconductor nanocrystals<br />

as fluorescent biological labels, Science, 281, 2013–2016, 1998.<br />

69. Sukhanova, A., Venteo, L., Devy, J., Artemyev, M., Oleinikov, V., Pluot, M., and Nabiev, I., Highly<br />

stable fluorescent nanocrystals as a novel class of labels for immunohistochemical analysis of<br />

paraffin embedded tissue sections, Laboratory Investigation, 82, 1259–1261, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Regulatory Perspective in Nanotechnology Products 155<br />

70. Sukhanova, A., Devy, J., Venteo, L., Kaplan, H., Artemyev, M., Oleinikov, V., Klinov, D., Pluot, M.,<br />

Cohen, J. K., and Nabiev, I., Biocompatible fluorescent nanocrystals for immunolabeling of<br />

membrane proteins and cells, Analytical Biochemistry, 324, 60–67, 2004.<br />

71. Lidke, D. S., Nagy, P., Heintzmann, R., Arndt-Jovin, D. J., Post, J. N., Freco, H. E., Jares-Erijman,<br />

E. A., and Jovin, T. M., Quantum dot ligands provide new insights into erbB/HER receptor mediated<br />

signal transduction, Nature Biotechnology, 22, 198–203, 2004.<br />

72. Akerman, M. E., Chan, W. C., Laakkonen, P., Bhatia, S. N., and Ruoslahti, E., Nanocrystal targeting<br />

in vivo, Proceedings of the National Academy of Sciences of the United States of America, 99,<br />

12617–12621, 2002.<br />

73. Kim, S., Lim, Y. T., Soltesz, E. G., De Grand, A. M., Lee, J., Nakayama, A., Parker, J. A.,<br />

Mihalijevic, T., Laurence, R. G., Dor, D. M., Cohn, L. H., Bawendi, M. G., and Frangioni, J. V.,<br />

Near-infrared fluorescent type II quantum dots for sentinel lymph node mapping, Nature Biotechnology,<br />

22, 93–97, 2003.<br />

74. Ballou, B., Lagerholm, B. C., Ernst, L. A., Bruchez, M. P., and Waggoner, A. S., Noninvasive<br />

imaging of quantum dots in mice, Bioconjugate Chemistry, 15, 79–86, 2004.<br />

75. Duncan, R., Nanomedicine gets clinical, Materials Today, 8 (8 Suppl. 1), 16–17, 2005.<br />

76. Yamawaki, H. and Iwai, N., Cytotoxicity of water-soluble fullerene in vascular endothelial cells,<br />

American Journal of Physiology. Cell Physiology, 290 (6), C1495–C1502, <strong>2006</strong>.<br />

77. Kamat, J. P., Devasagayam, T. P. A., Priyadarsini, K. I., and Mohan, H., Reactive oxygen species<br />

mediated membrane damage induced <strong>by</strong> fullerene derivatives and its possible biological implications,<br />

Toxicology, 155, 55–61, 2000.<br />

78. Sayes, C. M., Gobin, A. M., Ausman, K. D., Mendez, J., West, J. L., and Colvin, V. L., Nano-C60<br />

cytotoxicity is due to lipid peroxidation, Biomaterials, 26, 7587–7595, 2005.<br />

79. Dugan, L. L., Gabrielsen, J. K., Yu, S. P., Lin, T. S., and Choi, D. W., Buckminsterfullerenol free<br />

radical scavengers reduce excitotoxic and apoptotic death of cultured cortical neurons, Neurobiology<br />

of Disease, 3, 129–135, 1996.<br />

80. Hussain, S. M., Hess, K. L., Gearhart, J. M., Geiss, K. T., and Shlager, J. J., In vitro toxicity of<br />

nanoparticles in BRL 3A rat liver cells, Toxicology In Vitro, 19, 975–983, 2005.<br />

81. Jia, F., Wang, H., Yan, L., Wang, X., Pei, R., Yan, T., Zhao, Y., and Guo, X., Cytotoxicity of carbon<br />

nanomaterials: Single-wall nanotube, multi-wall nanotube, and fullerene, Environmental Science &<br />

Technology, 39, 1378–1383, 2005.<br />

82. Renwick, L. C., Donaldson, K., and Clouter, A., Impairment of alveolar macrophage phagocytosis<br />

<strong>by</strong> ultrafine particles, Toxicology and Applied Pharmacology, 172, 119–127, 2001.<br />

83. Moller, W., Hofer, T., Ziesenis, A., Karg, E., and Heyder, J., Ultrafine particles cause cytoskeletal<br />

dysfunctions in macrophages, Toxicology and Applied Pharmacology, 182, 197–207, 2002.<br />

84. Oberdorster, G., Toxicology of ultrafine particles: In vivo studies, Philosophical Transactions of the<br />

Royal Society of London, Series A, 358, 2719–2740, 2000.<br />

85. Brown, D. M., Wilson, M. R., Macnee, W., Stone, V., and Donaldson, K., Size-dependent proinflammatory<br />

effects of ultrafine polystyrene particles: A role for surface area and oxidative stress in<br />

the enhanced activity of ultrafines, Toxicology and Applied Pharmacology, 175, 191–199, 2001.<br />

86. Johnston, C. H., Finkelstein, J. N., Mercer, P., Corson, N., Gelein, R., and Oberdorster, G., Pulmonary<br />

effects induced <strong>by</strong> ultrafine PTFE particles, Toxicology and Applied Pharmacology, 168,<br />

208–215, 2000.<br />

87. Oberdorster, G., Oberdorster, E., and Oberdorster, J., Nanotoxicology: An emerging discipline<br />

evolving from studies of ultrafine particles, Environmental Health Perspectives, 113 (7), 823–839,<br />

2005.<br />

88. Holsapple, M. P. and Lehman-McKeeman, L. D., Forum series: Research strategies for safety<br />

evaluation of nanomaterials, Toxicological Sciences, 87 (2), 315, 2005.<br />

89. Oberdorster, E., Manufactured nanomaterials (fullerenes, C60) induce oxidative stress in the brain of<br />

juvenile largemouth bass, Environmental Health Perspectives, 112 (10), 1058–1062, 2004.<br />

90. Sullivan, D. C. and Ferrari, M., Nanotechnology and tumor imaging: Seizing an opportunity, Molecular<br />

Imaging, 3 (4), 364–369, 2004.<br />

91. Portney, N. G. and Ozkan, M., Nano-oncology: Drug delivery, imaging, sensing, Analytical and<br />

Bioanalytical Chemistry, 384, 620–630, <strong>2006</strong>.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


156<br />

Nanotechnology for Cancer Therapy<br />

92. Nanomedicine, an ESF-European Medical Research Councils, (EMRC) forward look report. http://<br />

www.esf.org/, 2005.<br />

93. El-Sayed, M., Ginski, M., Rhodes, C., and Ghandehari, H., Transepithelial transport or poly(amidoamine)<br />

dendrimers across Caco-2 cell monolayers, Journal of Controlled Release, 81, 355–365,<br />

2002.<br />

94. Jevprasesphant, R., Penny, J., Jalal, R., Attwood, D., McKeown, N. B., and D’Emanuele, A. D., The<br />

influence of surface modification on the cytotoxicity of PAMAM dendrimers, International Journal<br />

of Pharmaceutics, 252, 263–266, 2003.<br />

95. Lemarchand, C., Gref, R., Pasirani, C., Garcion, E., Petri, B., Muller, R., Costantini, D., and<br />

Couvreur, P., Influence of polysaccharide coating on the interactions of nanoparticles with biological<br />

systems, Biomaterials, 27, 108–118, <strong>2006</strong>.<br />

96. Kirchner, C., Liedl, T., Kudera, S., Pellegrino, T., Javier, A. M., Gaub, H. E., Stolzle, S., Fertig, N.,<br />

and Parak, W. J., Cytotoxicity of colloidal CdSe and CdSe/ZnS nanoparticles, Nano Letters, 5 (2),<br />

331–338, 2005.<br />

97. Neerman, M. F., Zhang, W., Parrish, A. R., and Simanek, E. E., In vitro and in vivo evaluation of a<br />

melamine dendrimer as a vehicle for drug delivery, International Journal of Pharmaceutics, 281,<br />

129–132, 2004.<br />

98. Nigavekar, S., Sung, L. Y., Llanes, M., El-Jawahri, A., Lawrence, R. S., Becker, C. W., Balogh, L.,<br />

and Khan, M. K., 3H Dendrimer nanoparticle organ/tumor distribution, Pharmaceutical Research,<br />

21 (3), 476–483, 2004.<br />

99. Malik, N., Wiwattanapatapee, R., Klopsch, R., Lorenz, K., Frey, H., Weener, J. W., Meijer, E. W.,<br />

Paulus, W., and Duncan, R., Relationship between structure and biocompatibility in vitro, and<br />

preliminary studies on the biodistribution of 125I-labeled polyamidoamine dendrimers in vivo,<br />

Journal of Controlled Release, 65, 133–148, 2000.<br />

100. Roberts, J. C., Bhalgat, M. K., and Zera, R. T., Preliminary biological evaluation of polyamidoamine<br />

(PAMAM) starburst dendrimers, Journal of Biomedical Materials Research, 30 (1), 53–65, 1996.<br />

101. Hardman, R., A toxicological review of quantum dots: Toxicity depends on physiochemical and<br />

environmental factors, Environmental Health Perspectives, 114 (2), 165–172, <strong>2006</strong>.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


9<br />

CONTENTS<br />

Polymeric Conjugates for<br />

Angiogenesis-Targeted Tumor<br />

Imaging and Therapy<br />

Amitava Mitra, Anjan Nan, Bruce R. Line, and<br />

Hamidreza Ghandehari<br />

9.1 Introduction ......................................................................................................................... 159<br />

9.2 Tumor Angiogenesis ........................................................................................................... 161<br />

9.3 Angiogenesis Markers ........................................................................................................ 162<br />

9.3.1 VEGF and VEGF Receptors .................................................................................. 162<br />

9.3.2 Aminopeptidases..................................................................................................... 162<br />

9.3.3 Matrix Metalloproteinases...................................................................................... 164<br />

9.3.4 Integrins .................................................................................................................. 166<br />

9.3.4.1 RGD Multimers ....................................................................................... 167<br />

9.4 Polymeric Conjugates for Angiogenesis Targeting............................................................ 168<br />

9.4.1 Polymer-Based Antiangiogenic Gene Therapy ..................................................... 169<br />

9.4.2 Polymer-Based Nuclear Imaging and Radiotherapy ............................................. 170<br />

9.5 Summary and Future Directions ......................................................................................... 175<br />

9.6 List of Abbreviations .......................................................................................................... 175<br />

Acknowledgment ......................................................................................................................... 177<br />

References .................................................................................................................................... 177<br />

9.1 INTRODUCTION<br />

Despite significant progress in the surgical treatment, radiotherapy, and chemotherapy of solid<br />

tumors (e.g., breast, colon, lung, and prostate), too few patients survive long term. This critical<br />

unmet need has been the long term focus of a wide range of therapies directed at the malignant<br />

cancer cell phenotype, its dysfunctional apoptotic signaling pathways, and survival adaptations. To<br />

enhance drug effectiveness and decrease non-specific toxicities, many of these agents have been<br />

designed to specifically target tumor cell antigens, receptors, and tumor microenvironmental<br />

markers.<br />

In contrast to direct targeting of specific tumor cell markers, targeting endothelial cells<br />

supporting tumor angiogenesis provides an alternative, broadly applicable method for cancer<br />

diagnosis and therapy. 1–5 All growing tumors require an augmented blood supply, so angiogenesis<br />

is a critical common denominator for therapeutic attack. Two distinct types of vessel targeted<br />

therapies have evolved. 6 The first general group, the anti-angiogenic agents, inhibit tumorinduced<br />

neovascularization <strong>by</strong> preventing proliferation, migration, and differentiation of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

159


160<br />

endothelial cells. 7–9 Anti-angiogenic drugs have been carried through substantial development to<br />

clinical trials. 7,10–12 Bevacuzimab (Avastin), an anti-vascular endothelial growth factor (VEGF)<br />

monoclonal antibody, has been approved for the treatment of colorectal cancer. 13 The second type,<br />

the antivascular agents, cause rapid and selective occlusion of existing tumor blood vessels, leading<br />

to tumor ischemia and extensive hemorrhagic necrosis. 5,14,15 Although still in preclinical development,<br />

antivascular-targeted therapeutics produce a characteristic pattern of necrosis after<br />

administration to mice and rats with solid tumors. 16,17 They cause a widespread central necrosis<br />

that can extend to as much as 95% of the tumor. 16–18<br />

Indeed, over the past 5–10 years, there has been a rapid development of small molecular<br />

therapies directed against neovascular angiogenesis. These molecules are typically smaller than<br />

2 kDa and show very low to absent immunogenicity, high target affinity, rapid targeting, and fast<br />

blood clearance. Small molecular agents have been directed at a number of angiogenesis-related<br />

targets and have been coupled to moieties to support scintigraphic, optical, and magnetic resonance<br />

imaging (MRI). 19,20 Other modifications have supported delivery of drugs, tumor enzymecleaved<br />

prodrugs, and therapeutic radionuclides. 21 This rich foundation provides the infrastructure<br />

upon which polymer-based conjugates have allowed further strides. Well-designed<br />

polymer conjugates demonstrate improved pharmacokinetics with prolonged blood residence,<br />

lower non-specific tissue penetration, and increased passive tumor localization because of<br />

enhanced permeability, and retention; i.e., the EPR effect. 22 Polymer conjugates bearing<br />

targeting ligands often demonstrate high tumor affinity via multivalent interactions. 23 Relative<br />

to smaller molecular platforms, they can also carry higher diagnostic and therapeutic agent<br />

payloads (Figure 9.1).<br />

(a)<br />

(c)<br />

(b)<br />

Biological Payload (Diagnostic or Therapeutic)<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 9.1 Schematic of angiogenesis-targeted conjugates. (a) Conjugate of cyclic RGD peptide with a<br />

therapeutic or a diagnostic agent. Chemotherapeutic or antiangiogenic drugs can be linked via biodegradable<br />

linkers, whereas imaging agents or radionuclides are generally attached via non-biodegradable linkers. (b)<br />

Particulate systems such as liposomes and nanoparticles are conjugated to cyclic RGD peptides via a nonbiodegradable<br />

linker, and the therapeutic or diagnostic agents are encapsulated. (c) Polymeric conjugates have<br />

cyclic RGD peptides linked onto the side chains via non-biodegradable linker, whereas biodegradable linkers<br />

can be used for drugs and non-biodegradable linkers for imaging agents or radionuclides.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 161<br />

Recent advances provide a clear indication that polyvalent, moderate molecular weight ligandpolymer<br />

conjugates will provide the platform to target angiogenesis and deliver a broad range of<br />

effective therapies. This chapter presents an overview of tumor angiogenesis, strategies for<br />

targeting angiogenic vasculature, and current developments in angiogenesis-targeted polymeric<br />

conjugates for diagnosis and therapy.<br />

9.2 TUMOR ANGIOGENESIS<br />

For tumors to grow and metastasize, they must secure an expanded blood supply <strong>by</strong> taking over<br />

existing blood vessels and promoting angiogenesis, i.e., the sprouting of new blood vessels from<br />

existing vessels. 2,3 Except in wound healing or in the female reproductive tissues, only 0.01% of<br />

normal endothelial cells are involved in angiogenesis. 24,25 In contrast, angiogenesis is both essential<br />

for tumor growth beyond 1–2 mm size and is highly specific for neoplasia. 26,27 Indeed, a highly<br />

significant statistical correlation has been shown between microvessel density and clinical stage,<br />

histopathology stage, and disease-specific survival. 28,29<br />

The induction of angiogenesis—the angiogenic switch—is an important early event in tumor<br />

progression and propagation. 30,31 The switch to an angiogenic phenotype is tightly regulated <strong>by</strong> a<br />

balance between endogenous anti-angiogenic 32–34 and pro-angiogenic 35,36 molecules. Studies in<br />

transgenic mice suggest that the switch occurs early in tumor development as a response to local<br />

stresses such as hypoxia and low pH. 37,38 Hypoxia increases levels of hypoxia-inducible factors<br />

(HIF) in tumor cells and in surrounding stromal cells. HIF then promotes the transcription of genes<br />

for VEGF and several other pro-angiogenic factors. 38,39<br />

During normal angiogenesis such as during reproduction, development, and wound healing,<br />

new vessels rapidly mature and stabilize. 2,4 By contrast, tumor blood vessels often fail to<br />

mature and are characterized <strong>by</strong> structures that are leaky, tortuous, and irregular in shape. 40<br />

The high levels of growth factors within a tumor support a continual state of vascular growth,<br />

regression, and regrowth. Many lack functional pericytes and are exceptionally permeable<br />

because of the presence of fenestrae, transcellular holes, and a lack of a complete basement<br />

membrane. 40–44 These ultrastructural changes lead to substantial leaks of tissue fluids into the<br />

tumor microenvironment and increased interstitial fluid pressure (IFP). Tumor IFP begins to<br />

increase as soon as the host vessels become leaky in response to angiogenic molecules such as<br />

VEGF. The importance of this transvascular flux is compounded <strong>by</strong> the impaired lymphatic<br />

system characteristic of cancer tissues. Hence, hyperpermeable angiogenic tumor vessels allow<br />

preferential extravasation of circulating macromolecules, and once in the interstitium, they are<br />

retained there <strong>by</strong> a lack of intratumoral lymphatic drainage. This enhanced permeability and<br />

retention (EPR) effect 22 results in intratumoral penetration and retention of macromolecules. It<br />

is one important factor explaining the delivery of macromolecular anti-cancer agents to the<br />

tumor microenvironment.<br />

In addition to the angiogenic factors that enhance the passive delivery of macromolecules via<br />

the EPR effect, angiogenic tumor vessels also display a variety of markers (Table 9.1) that can be<br />

the focus of active molecular targeting. These markers provide several targeting advantages<br />

relative to direct targeting of tumor cell. 45,46 First, endothelial cells are immediately accessible<br />

to intravenously delivered ligands. In contrast, the increased interstitial pressure associated with<br />

the tumor microenvironment can limit the access of macromolecular agents to tumor-associated<br />

targets. Second, endothelial cells are derived from normal, genetically stable host vessels. Hence,<br />

they present a target with characteristics that do not fluctuate unlike the genomic instability of<br />

most cancer tissues. Third, because large numbers of tumor cells rely on each blood vessel, antitumor<br />

effects will be amplified <strong>by</strong> vascular bed injury. Endothelial cell damage leads to vascular<br />

coagulation and downstream hypoxic death of approximately 100-fold more cancer cells. Fourth,<br />

endothelial targets are common to all solid tumors, whereas most other tumor cell associated<br />

markers are cancer phenotype specific.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


162<br />

TABLE 9.1<br />

Some Ligands and Their Target Receptors with Potential for Targeted Delivery to Tumor<br />

Angiogenesis<br />

9.3 ANGIOGENESIS MARKERS<br />

A relatively large number of endothelial cell surface proteins, ligands, and receptors can distinguish<br />

tumor angiogenic vessels from normal vasculature (Table 9.1). These markers on tumor endothelial<br />

cells have been identified <strong>by</strong> in vitro techniques (e.g., protein electrophoresis, mRNA-based serial<br />

analysis of gene expression, microarray analysis) 47 and in vivo techniques (e.g., biopanning of<br />

phage display libraries). 48,49 Many of these targets are the subject of reviews <strong>by</strong> Ruoshlathi and<br />

others. 43,44,47,50<br />

The identification of angiogenesis markers has spurred the development of technologies to<br />

enhance early cancer diagnosis, therapy planning, and therapeutic intervention. The development<br />

of diagnostic agents is mainly focused in the areas of nuclear scintigraphy, MRI, and near-infrared<br />

(NIR) imaging using matrix metalloproteinase (MMP) inhibitors and aVb3 integrin antagonists<br />

(Table 9.2). 19,20,51,52 Tumor vascular targeting peptides have been used to deliver chemotherapy<br />

(Table 9.3) given that they selectively target tumor angiogenesis 53,54 and are internalized once<br />

bound to cell surface receptors. 55,56 The most important classes of angiogenic targets and pertinent<br />

imaging and therapy studies are briefly discussed below.<br />

9.3.1 VEGF AND VEGF RECEPTORS<br />

VEGF has become one of the most important targets for antivascular therapy. It is an endothelial<br />

cell-specific mitogen that is a potent inducer of angiogenesis. It stimulates endothelial cell growth<br />

and acts through a family of closely related receptor tyrosine kinases, the most important of which<br />

is VEGFR-2. 36,57,58 VEGF receptor inhibitors have been shown to have anti-tumor effects. 59,60 The<br />

most clinically developed is a humanized anti-VEGF antibody (Bevacizumab) that has been<br />

approved <strong>by</strong> the U.S. Food and Drug Administration as a first-line therapy for metastatic colorectal<br />

cancer. 13<br />

9.3.2 AMINOPEPTIDASES<br />

Target Targeting Ligand a<br />

Reference<br />

aVb3 integrin RGD 53<br />

Aminopeptidase N (CD13) NGR 54<br />

Vascular endothelial growth factor Anti-VEGF antibody (Bevacizumab) 13<br />

aVb3 integrin Anti-aVb3 integrin antibody (Vitaxin) 193<br />

MMP-2 and MMP-9 HWGF 73<br />

Aminopeptidase A CPRECESIC 67<br />

Cell surface nucleolin F3 (N-terminal fragment of human high<br />

mobility group protein 2)<br />

194, 195<br />

Heparan sulfates CGKRK, CDTRL 196<br />

Unknown SMSIARL 197<br />

Kallikrein-9 CSRPRRSEC 196<br />

Platelet-derived growth factor receptor-b CRGRRST 198<br />

a Single amino acid codes for peptides (see list of abbreviations at end of chapter for details).<br />

Nanotechnology for Cancer Therapy<br />

Aminopeptidase N (APN) (also known as CD13) is a membrane, spanning 140 kD cell surface<br />

protein 61 that plays a role in tumor invasion. 62,63 APN has been successfully targeted using peptides<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 163<br />

TABLE 9.2<br />

Conjugates for Angiogenesis Targeted Imaging<br />

Compound Label Target Tumor Model Reference<br />

Nuclear imaging<br />

C(RGDfY) I-125 aVb3 integrin Murine osteosarcoma 117<br />

Gluco-RGD I-125 aVb3 integrin M21 human melanoma 120<br />

Galacto-RGD F-18 aVb3 integrin M21 human melanoma 121<br />

c(RGDf(N-Me)V) F-18 aVb3 integrin DLD-1 human colon<br />

adenocarcinoma<br />

199<br />

DOTA-c(RGDyK) Cu-64 aVb3 integrin MDA-MB-435 human<br />

mammary carcinoma<br />

118<br />

HYNIC-RGD4C Tc-99m aVb3 integrin Human renal adenocarcinoma 200<br />

FB-c(RGDyK)2 F-18 aVb3 integrin U87MG human glioblastoma 136<br />

DOTA-c(RGDyK)2 Cu-64 aVb3 integrin MDA-MB-435 human<br />

mammary carcinoma<br />

135<br />

HYNIC-E-c(RGDfK)2 Tc-99m aVb3 integrin OVCAR-3 human ovarian<br />

carcinoma<br />

107<br />

DOTA-E- c(RGDfK) 2 In-111 aVb3 integrin OVCAR-3 human ovarian<br />

carcinoma<br />

107<br />

DOTA-E{E- c(RGDfK)2}2 Cu-64 aVb3 integrin U87MG human glioblastoma 137<br />

FB-PEG-c(RGDyK) F-18 aVb3 integrin U87MG human glioblastoma 140<br />

DOTA-PEG-c(RGDyK) Cu-64 aVb3 integrin U87MG human glioblastoma 141<br />

DOTA-PEG-E-c(RGDyK)2 Cu-64 aVb3 integrin Human female lung<br />

adenocarcinoma<br />

201<br />

HPMA copolymer-RGD4C Tc-99m aVb3 integrin DU145 & PC-3 human prostate<br />

carcinoma<br />

23, 178<br />

HPMA copolymer-RGDfK In-111 aVb3 integrin Murine lewis lung carcinoma 181<br />

CGS27023A F-18 MMP Murine ehrlich breast tumor 85, 87<br />

SAV03M F-18 MMP Murine ehrlich breast tumor 87<br />

D-Tyr-HWGF I-125 MMP-2, MMP-9 Murine lewis lung carcinoma 83<br />

DOTA-HWGF<br />

Magnetic resonance imaging<br />

Cu-64 MMP-2, MMP-9 MDA-MB-435 human<br />

mammary carcinoma<br />

90, 91<br />

Anti-aVb3 antibody (LM609)liposome<br />

Gadolinium aVb3 integrin Rabbit V2 carcinoma 126<br />

Anti-aVb3 antibody (DM101)- Gadolinium aVb3 integrin Rabbit corneal micropocket 127<br />

nanoparticle<br />

model<br />

aVb3 peptidomimetic antagonistnanoparticle<br />

Optical imaging<br />

Gadolinium aVb3 integrin Rabbit V2 carcinoma 110<br />

Cy5.5-c(RGDyK) Cy5.5(NIR dye) aVb3 integrin U87MG human glioblastoma 111<br />

Cy5.5- E- c(RGDyK) 2 Cy5.5 aVb3 integrin U87MG human glioblastoma 129<br />

Cy5.5-E{E- c(RGDfK)2}2 Cy5.5 aVb3 integrin U87MG human glioblastoma 129<br />

Cy5.5-GPLGVRGK-PEG-PLL Cy5.5 MMP-2 HT1080 human fibrosarcoma 94<br />

NIRQ820-GVPLSLTMGC-Cy5.5 Cy5.5 MMP-7 HT1080 human fibrosarcoma 97<br />

containing an Asn-Gly-Arg (NGR) motif. 54 NGR peptides have been used to deliver chemotherapeutic<br />

agents 64 and pro-apoptotic peptides 65 to tumor cells. In a murine breast carcinoma xenograft<br />

model, a NGR-doxorubicin conjugate significantly decreased metastasis to lymph nodes and<br />

substantially increased survival compared to free doxorubicin. In addition, the conjugates were<br />

less toxic to the liver and heart as compared to the free drug. 64<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


164<br />

Aminopeptidase A (APA) is a homodimeric membrane, spanning cell surface protein that is<br />

upregulated in the pericytes of tumor blood vessels. 66 Phage display studies have identified an APA<br />

binding peptide (CPRECESIC) that inhibits enzymatic activity of APA, suppresses both migration<br />

and proliferation of endothelial cells, and inhibits tumor growth. 67<br />

9.3.3 MATRIX METALLOPROTEINASES<br />

MMPs are a family of over 21 zinc-dependent neutral endopeptidases that play an important role in<br />

tumor angiogenesis, tissue remodeling, and cell migration. 68–70 The two most important MMPs in<br />

cancer progression are MMP-2 (gelatinase A) and MMP-9 (gelatinase B). 71,72 In cancer, levels of<br />

these MMPs are abnormally elevated, enabling cancer cells to degrade the extracellular matrix<br />

(ECM), invade the vascular basement membrane, and metastasize to distant sites. 70<br />

MMP targeting for tumor imaging or therapy can be achieved using small molecule inhibitors<br />

of MMPs, peptide sequences that target MMP, tissue inhibitors of MMPs, and MMP peptide<br />

substrates that are cleaved at the tumor site. For example, cyclic peptides containing a HWGF<br />

motif can specifically inhibit MMP-2 and MMP-9 activities and prevent both tumor growth and<br />

metastasis. 73 Several small molecule MMP inhibitors such as the sulfonamide–hydroxamate<br />

compound CGS 27023A are currently in clinical trials. 69,74–76<br />

The protease functionality of MMP provides additional opportunities useful in enzyme prodrug<br />

therapy (Table 9.3). For example, drugs such as doxorubicin, auristatins, and duocarmycin may be<br />

conjugated to an acetyl-PLGL peptide sequence, MMP substrate, such that when incubated with<br />

MMP-2 or MMP-9, cleavage at the GL bond liberates a leucyl drug. Kline et al. showed that such a<br />

strategy produced no other cleavage intermediates and that proteolysis products were equipotent<br />

with the parent drugs. 77<br />

MMP-substrate-doxorubicin conjugates are found to be preferentially metabolized in murine<br />

fibrosarcoma xenografts relative to plasma and cardiac tissues. 78 There was a 10-fold increase in<br />

TABLE 9.3<br />

Conjugates for Angiogenesis Targeted Drug Delivery<br />

Nanotechnology for Cancer Therapy<br />

Compound Drug Target Tumor Model Reference<br />

NGR-doxorubicin Doxorubicin Aminopeptidase N MDA-MB-435 human mammary<br />

carcinoma<br />

64<br />

mPEG-GPLGVdoxorubicin<br />

Doxorubicin MMP-2, MMP-9 Murine lewis lung carcinoma 79<br />

Albumin- GPLGIAGQdoxorubicin<br />

Doxorubicin MMP-2, MMP-9 A375 human melanoma 82<br />

PLGL-doxorubicin Doxorubicin MMP-2, MMP-9 HT1080 human fibrosarcoma 78<br />

RGD4C-doxorubicin Doxorubicin aVb3 integrin MDA-MB-435 human mammary<br />

carcinoma<br />

64<br />

RGD4C-doxorubicin Doxorubicin aVb3 integrin MH134 murine hepatoma 132<br />

RGD4C-AFK-doxorubicin Doxorubicin aVb3 integrin HT1080 human fibrosarcoma 130, 131<br />

E-c(RGDyK) 2-paclitaxel Paclitaxel aVb3 integrin MDA-MB-435 human mammary<br />

carcinoma<br />

202<br />

PTK787-albumin-PEG- PTK787 (kinase aVb3 integrin Human umbilical vein endothelial cells 21<br />

c(RGDfK)<br />

inhibitor)<br />

c(RGDfC)-PEG-liposomes 5-FU aVb3 integrin Murine B16F10 melanoma 185<br />

RGD-PEG-liposomes Doxorubicin aVb3 integrin Murine B16F10 melanoma 203<br />

c(RGDfC)-PEG- Doxorubicin aVb3 integrin Cl-66 murine metastatic mammary 186<br />

nanoparticles<br />

tumor carcinoma<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 165<br />

doxorubicin tumor-to-heart ratio and a greater effectiveness than free doxorubicin at reducing<br />

tumor growth. In particular, the conjugate cured 8 of 10 mice at levels below the maximum<br />

tolerated dose, whereas doxorubicin cured 2 of 20 mice at its maximum tolerated dose. Furthermore,<br />

mice treated with the conjugate had no detectable change in body weight or circulating<br />

reticulocytes. 78<br />

Other studies have focused on molecular modifications to modify in vivo biodistribution<br />

kinetics. For example, MMP substrate peptide–doxorubicin conjugates have been coupled to polyethylene<br />

glycol (PEG) to increase blood circulation time and to reduce non-specific cytotoxicity. 79<br />

The anti-cancer drug conjugate, mPEG–GPLGV–DOX, was synthesized <strong>by</strong> conjugating doxorubicin<br />

with GPLGV peptide (a substrate for MMP-2 and MMP-9) and PEG methyl ether (mPEG).<br />

In vivo experiments showed that a 50 mg/kg dose of mPEG–GPLGV–DOX had similar therapeutic<br />

effectiveness as a 10 mg/kg dose of doxorubicin, but it was associated with lower toxicity and<br />

prolonged life span relative to free drug. 79<br />

An interesting strategy to deliver drug therapy <strong>by</strong> EPR was investigated using a maleimide–<br />

doxorubicin conjugate of an octapeptide MMP substrate (GPLGIAGQ). 80 The water-soluble<br />

substrate was designed to form an in situ complex with an albumin thiol (cysteine-34). 81 The<br />

conjugate was efficiently cleaved <strong>by</strong> activated MMP-2 and MMP-9, liberating a doxorubicin<br />

tetrapeptide that showed antiproliferative activity in vitro in a murine renal cell carcinoma.<br />

When tested in a murine model of a human melanoma xenograft, the doxorubicin–octapeptide<br />

conjugate showed complete binding with albumin within 5 min and a 16 h stability half-life. 82 The<br />

maximum tolerated dose of the conjugate was 4-fold greater than free doxorubicin. Furthermore,<br />

the conjugate showed superior anti-tumor effects as compared to free doxorubicin with duration of<br />

remission for up to 40 days and almost a 4-fold reduction in tumor size as compared to free<br />

doxorubicin at the end of the 50 day trial period.<br />

Although imaging MMP expression is a relatively new area of research, a number of radiolabeled<br />

imaging agents have been tested in animal models (Table 9.2). 19,20 I-125 labeled cyclic<br />

decapeptides containing HWGF have shown relatively disappointing results in mice bearing Lewis<br />

lung carcinoma. 83 There was low to moderate uptake in tumor, significant tumor washout overtime,<br />

high concentrations in the liver and kidney, and low metabolic stability of the iodo-tyrosine<br />

complex. 83 More promising results have been obtained with 18 F and 11 C labeled MMP inhibitors<br />

developed for positron emission tomography (PET) imaging. 84–86 In vitro studies showed no loss of<br />

MMP inhibitory activity of the radiolabeled compound as compared to the parent compound. 84,87<br />

Dynamic microPET images of tumor-bearing mice showed substantially higher tumor uptake than<br />

other background organs. 88,89 The cyclic HWGF peptide has been conjugated with a 1,4,7,10-tetraazacylcododecane-N,N<br />

0 ,N 00 ,N 000 -tetraacetic acid (DOTA) chelate to enable 64 Cu PET imaging. 90,91<br />

Studies of the 64 Cu labeled peptide in a murine breast cancer model showed improved pharmacokinetics,<br />

tumor accumulation, and metabolic stability relative to a comparable 125 I<br />

radiolabeled peptide.<br />

111 In-diethylenetriaminepentaacetic acid (DTPA)-N-TIMP-2, a radiolabeled high affinity<br />

endogenous tissue inhibitor of MMP, has been found to have similar activity as unmodified<br />

N-TIMP-2. 92 Unfortunately, 111 In-DTPA-N-TIMP-2 yielded disappointing results with insignificant<br />

tumor uptake in clinical studies in patients with Kaposi’s sarcoma. 93<br />

Optical imaging methods to assess MMP activity has been achieved using a NIR fluorescence<br />

probe attached to a MMP-2 substrate. 94–96 The MMP-2 imaging probe consists of three elements: a<br />

quenched NIR fluorochrome (Cy5.5), a MMP-2 peptide substrate (GPLGVRGK), and a PEG-poly-<br />

L-lysine (PEG–PLL) graft copolymer. 94 The conjugate is designed such that the peptides are<br />

cleaved <strong>by</strong> MMP-2 at the tumor site, resulting in several hundred percent increases in the NIR<br />

fluorescence, allowing visualization of the MMP-2 in human fibrosarcoma tumor. Next MMP-2<br />

inhibition was imaged after treatment with prinomastat where the treated tumor showed significantly<br />

lower NIR signal as compared to the untreated tumors. 94 A similar NIR fluorescencebased<br />

probe has been designed for MMP-7 that is over expressed in colon, breast, and pancreatic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


166<br />

cancer. 97 In tumor extract containing MMP-7 in vitro, the fluorescence signal from the probe was<br />

enhanced 7-fold after enzymatic degradation. This probe could also be used to image tumorassociated<br />

enzyme activity and also inhibition of MMP-7 in vivo.<br />

9.3.4 INTEGRINS<br />

Nanotechnology for Cancer Therapy<br />

Integrins are a diverse group of heterodimeric cell surface receptors for ECM molecules. 98,99 This<br />

group consists of at least 25 integrins through the pairing 18 a and 8 b subunits. They mediate cell<br />

adhesion, 100 help in cell migration during metastasis, 101 and regulate cell growth. 102 a Vb 3, a Vb 5,<br />

and a 5b 1 integrins are upregulated in angiogenic endothelial cells and are essential for angiogenesis.<br />

103–106<br />

High affinity aVb3 (and aVb5) selective ligands containing the tripeptide Arg-Gly-Asp (RGD)<br />

have been identified <strong>by</strong> phage display studies 53 and have been used to deliver chemotherapeutic<br />

agents, 64 pro-apoptotic peptides, 65 and radiotherapy 107,108 to tumor tissues. RGD peptides and<br />

RGD mimetics 109 have been used to detect cancer when coupled to image contrast agents designed<br />

for MRI, 110 gamma scintigraphy, 19 PET, 19,51 and NIR fluorescence imaging. 111<br />

A large number of radiolabeled tracers based on the RGD tripeptide sequence have been<br />

developed (Table 9.2). 19,51,52 Structure activity relationship investigations of the cyclic pentapeptide,<br />

cyclo-(Arg 1 -Gly 2 -Asp 3 -D-Phe 4 -Val 5 ), 112–114 revealed that a hydrophobic amino acid in<br />

position 4 is essential for high aVb3 affinity and that the amino acid in position 5 could be modified<br />

without loss of activity. 115 A tyrosine residue at position 4 or 5 provides a site to radio-iodinate the<br />

peptide, and lysine residue introduced at position 5 enables 18 F labeling. 116,117 Besides radiohalogenation,<br />

cyclic RGD peptides can be labeled with radio-metals <strong>by</strong> introducing a lysine residue at<br />

the terminal end of the peptide and conjugating that with a chelator such as DOTA, 118 DTPA, 119<br />

hydrazinonicotinamide (HYNIC), 107 or N-u-bis(2-pyridylmethyl)-L-lysine (DPK). 23 Other radiometals<br />

such as yttrium-90 ( 90 Y) may provide a means to deliver molecularly targeted radiotherapy.<br />

For example, in a murine ovarian carcinoma xenograft model, a single injection of a 90 Y labeled,<br />

dimeric-RGD peptide-DOTA conjugate at its maximum tolerated dose (37 MBq) caused significant<br />

tumor growth delay and increased median survival time. 107<br />

Unfortunately, many of these tracers are lipophilic and are both concentrated and excreted <strong>by</strong><br />

the liver. 117 Non-tumor bearing organs that retain the tracers may receive an unacceptably high<br />

radiation dose. To improve their pharmacokinetics, glucose or a galactose-based sugar amino acid<br />

has been introduced into the pentapeptide architecture. The resulting iodo-gluco-RGD 120 or [ 18 F]galacto-RGD<br />

121,122 shows reduced concentration in the liver and increased accumulation in<br />

tumor. In other attempts to improve pharmacokinetics, cyclic-RGD peptides have been conjugated<br />

to tetrapeptides of D-amino acids. 123 The pharmacokinetics and tumor accumulation of the<br />

resulting 18 F labeled conjugates containing D-asparatic acid are found to be comparable to [ 18 F]galacto-RGD.<br />

123<br />

Non-invasive imaging of angiogenesis markers should be important, both in therapy planning<br />

and in monitoring anti-angiogenic therapy. The first human images of aVb3 expression have been<br />

obtained with [ 18 F]-galacto-RGD in patients with malignant melanoma and sarcoma. 124 These<br />

biodistribution studies demonstrated highly favorable pharmacokinetic profiles with specific<br />

receptor binding, rapid blood clearance primarily through the kidney, increasing tumor-to-background<br />

(T/B) ratios with time, and almost 4-fold higher tumor distribution volume as compared to<br />

muscle. 125 Tumor aVb3 expression was imaged with good contrast in most parts of the body<br />

with the exception of the region near the urogenital tract, the tracer excretion route.<br />

Detection of tumor angiogenesis <strong>by</strong> aVb3 targeted MRI has been achieved using liposomes or<br />

nanoparticles encapsulating gadolinium ion that was derivatized with anti-aVb3 antibody<br />

LM609, 126 antibody DM101 127 or RGD mimetics. 110,128 This approach provided detailed images<br />

of the tumor and delineation of the tumor from surrounding tissues, and it highlighted angiogenic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 167<br />

vasculature. This enhanced detection of tumors may be helpful in directing therapeutic<br />

interventions.<br />

Several optical angiogenesis imaging methods using cyclic RGD peptides have been described.<br />

One such method utilized a NIR dye Cy5.5 conjugated to RGD to visualize glioblastoma xenograft<br />

in nude mice. 111 The tumor could be visualized with high contrast to the background organs as early<br />

as 30 min post-injection. This study showed the feasibility of in vivo optical imaging of aVb3<br />

expressions. In a subsequent study, RGD mono-, di- and tetramer conjugated to Cy5.5 were<br />

compared in a glioblastoma xenograft model. 129 The tumor xenograft could be visualized with<br />

all three probes 30 min post-injection. The tumor uptake was in the order tetramerOdimerO<br />

monomer; however, the increase in the T/B ratios for the multimers was modest. For example,<br />

the dimer and the tetramer showed lower tumor-to-kidney ratios than the monomer because of<br />

higher net positive charge. Therefore, only modest enhancement of therapeutic index is possible <strong>by</strong><br />

multimerization of RGD peptides and further optimization of these probes is needed.<br />

RGD-targeted prodrugs have been tested as a means to improve chemotherapy toxicity profiles<br />

(Table 9.3). One strategy utilizes an enzymatically cleavable tripeptide sequence (D-Ala-Phe-Lys)<br />

recognized <strong>by</strong> the tumor-associated protease, plasmin, as a linker between RGD4C and doxorubicin.<br />

130 Upon incubation with plasmin in vitro, doxorubicin is released and regains its cytotoxicity<br />

for endothelial and fibrosarcoma cells. Toxicity studies in BALB/c and nude mice showed significantly<br />

higher weight loss and mortality for free doxorubicin relative to an equimolar dose of the<br />

prodrug. 131 Furthermore, in vivo evaluation in mouse breast cancer and human adenocarcinoma<br />

models showed that prodrug anti-tumor efficacy was associated with strong inhibition of angiogenesis.<br />

131 A similar RGD4C-doxorubicin conjugate was tested in an a Vb 3 negative murine hepatoma<br />

model and was found to have superior anti-tumor effectiveness as compared to free doxorubicin. 132<br />

9.3.4.1 RGD Multimers<br />

As experience with RGD-targeted compounds has evolved, efforts to enhance effectiveness have<br />

focused on improving tumor uptake and reducing non-specific localization in normal tissues. To<br />

enhance tumor targeting, RGD multimers have been studied as means to increase receptor binding<br />

affinity through polyvalency. A dimeric RGD peptide labeled with 111 In, 107 99m Tc, 133,134 64 Cu, 135<br />

and 18 F 136 have all shown higher receptor binding affinity and significantly higher tumor accumulation<br />

than the RGD monomer. Similarly, tetrameric RGD peptides have shown almost 3-fold and<br />

10-fold higher integrin affinity than the dimer and monomer, respectively. 129,137 Biodistribution<br />

studies in M21 melanoma xenograft-bearing mice indicated that tumor uptake and T/B ratios<br />

increased as monomer!dimer!tetramer. 138 However, more recent biodistribution studies in<br />

human glioblastoma-bearing nude mice revealed that the higher tumor uptake of the tetramer as<br />

compared to dimer was offset <strong>by</strong> increased liver and kidney uptake. 129,137 This significantly<br />

lowered the T/B ratios of the multimers, substantially reducing the therapeutic index.<br />

To address the unfavorable pharmacokinetics of these cyclic peptides, investigators have<br />

studied their conjugation to hydrophilic polymers such as PEG. A comparative biodistribution<br />

study showed that free peptide had a more rapid tumor washout than a pegylated cyclic RGD<br />

conjugate. In contrast, the PEG–RGD conjugate showed a gradual increase in both tumor accumulation<br />

and tumor-to-blood ratios. 139 Similarly, studies of 18 F 140 and 64 Cu 141 radiolabeled pegylated<br />

RGD have yielded high quality microPET images of glioblastoma xenografts with higher T/B ratios<br />

than the non-pegylated peptide.<br />

As might be expected, studies combining multivalency and pegylation have shown significantly<br />

improved tumor retention and tumor-to-normal tissue distribution ratios. 142 The tumor uptake of<br />

pegylated RGD multimers increased <strong>by</strong> about 3-fold from monomer to tetramer. Tetramer tumorto-liver<br />

ratios were 20- and 3-fold higher than the monomer and dimer, respectively. 142<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


168<br />

9.4 POLYMERIC CONJUGATES FOR ANGIOGENESIS TARGETING<br />

Polymeric conjugates have been used for targeted delivery of drugs to tumor sites 143–146 because<br />

the attachment of drugs to water soluble polymers increases their aqueous solubility, reduces side<br />

effect, and overcomes multi-drug resistance; the large size of the conjugates increases the blood<br />

half-life and significantly alters the drug properties and pharmacokinetics; the conjugates can be<br />

tailor-made (i.e., side-chain content, molecular weight, charge, etc.) for specific targeting and<br />

delivery needs; they can be designed to passively (EPR) or actively target tumor sites; and sitespecific<br />

drug release can be achieved <strong>by</strong> designing biodegradable spacers that can be enzymatically<br />

cleaved or that are pH sensitive. These advantages have led to the development of a wide range of<br />

polymer-anti-cancer drug conjugates, some of which are currently in clinical trials. 145<br />

N-(2-hydroxypropyl) methacrylamide (HPMA) copolymers have shown promise as drug<br />

carriers. 144,145,147,148 HPMA copolymers have been employed to modify the in vivo biodistribution<br />

of chemotherapeutic agents and enzymes. The advantages of HPMA copolymers over other watersoluble<br />

polymers is that they can be tailor-made with simple chemical modifications to regulate<br />

drug and targeting moiety content for biorecognition, internalization, or subcellular trafficking<br />

depending on specific therapeutic needs (Figure 9.2). 144,149,150 The overall molecular weight of<br />

HPMA copolymers is determined <strong>by</strong> the polymerization conditions, particularly the concentration<br />

of initiator and chain transfer agents. 151 Various side chain moieties (isotope chelators, targeting<br />

moieties, and drugs) (Figure 9.2) may be directly linked to the polymer chain via a biodegradable or<br />

non-biodegradable spacer. 143,144<br />

Polymer-based delivery systems have been used as carriers for passive and active targeting of<br />

drugs in the treatment of various diseases and as novel imaging agents. 145,152 Without a specific<br />

targeting ligand, moderate-sized (O30 kD) polymers can passively (via EPR) accumulate in tumor<br />

tissues. 22,153 The EPR effect has been used to deliver macromolecular bioactive agents to solid<br />

tumors, including anti-angiogenic drugs. 154,155<br />

There are a number of important differences between small molecular weight drugs and<br />

polymeric conjugates of these small molecules. The advantage of polymer-based delivery<br />

1 2 3 4 5<br />

H CH 3<br />

2<br />

( C C )<br />

x<br />

C O<br />

CH 3<br />

H 2<br />

( C C )<br />

y<br />

C O<br />

H CH 3<br />

2<br />

( C C )<br />

m<br />

C O<br />

H<br />

CH 3<br />

2<br />

(C C )<br />

n<br />

C O<br />

NH<br />

NH O<br />

NH<br />

NH<br />

CH 2 HC C NH 2 CH 2<br />

CH 2<br />

HC OH CH 2<br />

C O<br />

C O<br />

CH 3<br />

NH<br />

NH<br />

OH<br />

CH 2<br />

C O<br />

NH HOOC<br />

CH 2<br />

C O<br />

H<br />

C NH<br />

Peptide CH 2<br />

RGD4C, RGDfK<br />

CH 2<br />

CH 2<br />

CH 2<br />

H N<br />

2C 2C CH 2<br />

125 I, 123 I, 124 I, 131 I<br />

N<br />

188 Re, 99m Tc<br />

N<br />

H CH<br />

2<br />

3<br />

( C C )<br />

z<br />

C O<br />

NH<br />

CH 2<br />

CH 2<br />

CH 2<br />

NH<br />

C S<br />

NH<br />

COOH<br />

N<br />

COOH<br />

N<br />

COOH<br />

N<br />

COOH<br />

COOH<br />

Nanotechnology for Cancer Therapy<br />

177 Lu, 90 Y, 213 Bi, 210 Po<br />

FIGURE 9.2 Structure of water-soluble HPMA copolymers for angiogenesis targeted delivery of radionuclides<br />

for imaging and therapy. The conjugate can contain side-chains of different chelating comonomers<br />

for a wide range of radionuclides for use as a single agent for both diagnosis and therapy. 1-HPMA, 2-MA-Tyr,<br />

3-MA-GG-RGD4C, 4-MA-GG-DPK, and 5-APMA-CHX-A 00 -DTPA (see list of abbreviations for details).<br />

(From Mitra, A., et al., Nucl. Med Biol., 33, 43, <strong>2006</strong>. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 169<br />

systems stem from decreased extravasation in normal tissues because of the large molecular weight<br />

of the conjugates. 156 Low extravasation of polymer conjugates in normal tissues generally results in<br />

reduced systemic toxicity. In this regard, the predominant liver 117 and kidney 120 uptake of small<br />

RGD peptides has been identified as a significant disadvantage of targeting tumor angiogenesis with<br />

small peptides.<br />

The polymeric backbone may also provide a platform to support multivalent targeting ligands<br />

(Figure 9.1). As previously discussed, conjugation of a targeting peptide onto a polymer backbone<br />

significantly enhances the tumor to normal tissue uptake in comparison to the peptide itself. This is<br />

likely attributable to an increased target affinity because of the multivalency of the targeting moiety<br />

on the polymer backbone, 157 a combination of active targeting and passive EPR effect of the<br />

macromolecular conjugate, 158 and a decreased extravasation in normal tissues as a result of the<br />

large molecular weight of the conjugates. 156<br />

The strategy of targeting the tumor angiogenic endothelium using polymeric conjugates is<br />

particularly attractive where angiogenesis inhibitors show substantial toxicity at the effective<br />

dose, e.g., the fumagillin analogue TNP-470 shows extreme neurotoxicity at this level. 159 Conjugation<br />

of TNP-470 to a water-soluble polymer via an enzymatically degradable spacer allows<br />

passive localization and drug release at the tumor site and prevents blood–brain barrier<br />

penetration. 154<br />

9.4.1 POLYMER-BASED ANTI-ANGIOGENIC GENE THERAPY<br />

Anti-angiogenic gene therapy strategies against several human tumors have shown encouraging<br />

results in animal models. 160,161 A number of strategies to deliver therapeutic genes to angiogenic<br />

endothelial cells using RGD guided viral 162,163 and non-viral systems (cationic polymers, 164<br />

cationic lipids, 165 and cationic peptides 166 ) have been investigated. The cationic polymer polyethylenimine<br />

(PEI) has been extensively used as a non-viral gene carrier. 167 PEI has also been<br />

investigated as an RGD guided transfection system for angiogenic endothelial cells (Table 9.4).<br />

Broadly, these systems can be divided into two classes: those that are direct conjugates of<br />

PEI–RGD 168,169 and those where RGD is conjugated onto PEI via a hydrophilic PEG spacer<br />

(PEI–PEG–RGD). 164,168,170–172 Because of their reduced non-specific interaction with normal<br />

tissues, the charge-shielded pegylated systems have shown more efficient targeting and gene<br />

transfection. 171,172<br />

The PEI–PEG–RGD conjugate architecture has been used to deliver siRNA as a means to<br />

inhibit VEGFR-2 expression. 170 Specific tumor accumulation was observed in a murine neuroblastoma<br />

model, and there was lower non-specific uptake in lung and liver as compared to aqueous<br />

siRNA and non-targeted conjugate. With a 40 mg treatment dose repeated every three days, there<br />

TABLE 9.4<br />

Conjugates for Angiogenesis Targeted Gene Therapy<br />

Compound Gene Target Tumor Model Reference<br />

c(RGD)-PEG-PEI siRNA aVb3 integrin N2A neuroblastoma 170<br />

RGD4C-PEG-PEI sFlt-1 aVb3 integrin Human skin capillary endothelial cells 164<br />

Pronectin F Cw<br />

LacZ aVb3 integrin Murine meth-AR-1 fibrosarcoma 173<br />

RGDC-PEG-PEI Luciferase aVb3 integrin Mewo human melanoma 168<br />

RGD-PEI Plasmid DNA aVb3 integrin Human cervical carcinoma 169<br />

RGD-PEG-PEI DNA aVb3 integrin Not reported 172<br />

RGD-lipid nanoparticles Raf-1 aVb3 integrin CT26 human colon carcinoma 165<br />

RGD4C-PEG-cholesterol liposome CAT aVb3 integrin PO2 human colon carcinoma 204<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


170<br />

was significant inhibition of tumor angiogenesis, reduced tumor growth rate, and a marked<br />

reduction of peritumoral vascularization. 170 In another study, Kim et al. reported an anti-angiogenesis<br />

strategy to block VEGF receptor function and there<strong>by</strong> inhibit endothelial cell proliferation. 164<br />

The authors conjugated PEI–PEG–RGD and a sFlt-1 gene that encoded a soluble fragment of Flt-1<br />

(VEGF receptor antagonist) and found that this non-viral gene delivery system enabled stable<br />

expression of sFlt-1 <strong>by</strong> endothelial cells. The expressed soluble Flt-1 fragment, coupled to exogenous<br />

VEGF, blocked Flt-1 receptor binding and inhibited cultured endothelial cell proliferation. 164<br />

Using another approach, Hosseinkhani and Tabata investigated the feasibility of tumor targeting<br />

using a non-viral gene carrier synthesized from a genetically engineered silk-like polymer<br />

containing repeated RGD sequences(Pronectin FC). 173 When cationized, Pronectin FC-plasmid<br />

DNA complexes with or without pegylation were intravenously injected into mice carrying a<br />

subcutaneous Meth-AR-1 fibrosarcoma mass, and the level of gene expression in the tumor was<br />

significantly higher for the PEG–DNA complex relative to non-pegylated complexes and free<br />

plasmid DNA.<br />

Although anti-angiogenic gene therapy is still in its infancy, the studies beginning to appear in<br />

the literature suggest the prospects of development of new polymer-based gene delivery systems<br />

with improved transfection efficiency and reduced toxicity. It appears that many of the targeting and<br />

biodistribution lessons learned in studies of polymer-based drug delivery to sites of angiogenesis<br />

may be applied to enhance the success of polymer based gene delivery as well.<br />

9.4.2 POLYMER-BASED NUCLEAR IMAGING AND RADIOTHERAPY<br />

With the appropriate delivery system, radioisotopes have a significant advantage over other therapy<br />

agents, namely, the emission of energy that can kill at a distance from the point of radioisotope<br />

localization. This diameter of effectiveness helps to overcome the problem of tumor heterogeneity<br />

because, unlike other molecular therapy (cell toxins, chemotherapy, etc.), not all tumor cells need to<br />

take up the radioisotope to eradicate a tumor. There are also physical characteristics (type of<br />

particle emission, emission energy, half-life) of different radioisotopes that may be selected to<br />

enhance therapeutic effectiveness. 174 For example, different isotopes deliver beta particulate ionization<br />

over millimeters ( 131 I) to centimeters ( 90 Y). Long-lived isotopes such as 131 I that remain<br />

within the tumor target may provide extended radiation exposure and high radiation dose,<br />

especially if there is progressive renal clearance and high target to non-target ratios.<br />

RGD peptides labeled with therapeutically relevant isotopes such as b-particle emitters have<br />

been investigated as potential angiogenesis targeted radiotherapy (Table 9.5). The chelation conditions<br />

for 90 Y and lutetium-177 ( 177 Lu) labeled RGD have revealed that time, temperature, pH,<br />

presence of trace metal contaminants, and stoichiometric ratio of chelator to isotope all have<br />

significant effects on the rate of chelation and radiolabeling efficiency. 175 A major challenge in<br />

development of therapeutic radiopharmaceuticals is radiolytic degradation of radiolabeled products<br />

because of production of free radicals in the presence of a large amount of high energy<br />

b-particles. 176 A study on the stability of 90 Y labeled, dimeric RGD peptide showed that presence<br />

TABLE 9.5<br />

Conjugates for Angiogenesis Targeted Radiotherapy<br />

Nanotechnology for Cancer Therapy<br />

Compound Radiolabel Target Tumor Model Reference<br />

DOTA-E- c(RGDfK)2 Y-90 aVb3 integrin OVCAR-3 human ovarian carcinoma 107, 205<br />

DTPA- Tyr3-octreotate-c(RGDyD) In-111 aVb3 integrin CA20948 & AR42J rat<br />

pancreatic tumor<br />

206,207<br />

HPMA copolymer-RGD4C Y-90 aVb3 integrin DU145 human prostate carcinoma 108<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 171<br />

of free radical scavengers such as gentisic acid (GA) and ascorbic acid (AA), together with storage<br />

at low temperature (K788C), stabilized the labeled compound for at least two half-lives of 90 Y,<br />

even at a high specific activity. 177<br />

To develop an HPMA copolymer, RGD-based targeted angiogenesis imaging and therapy<br />

agent, the biodistribution and tumor accumulation of HPMA copolymer-RGD4C conjugate as<br />

compared to a control conjugate containing the tripeptide Arg-Gly-Glu (HPMA copolymer-<br />

RGE4C conjugate) has been studied. 178 Figure 9.3 shows the scintigraphic image of SCID mice<br />

bearing prostate tumor xenografts 24 h post-injection of the conjugates. The HPMA copolymer-<br />

RGD4C conjugate shows greater tumor accumulation than the control. The tumor uptake of HPMA<br />

copolymer-RGD4C conjugate (4.6G1.8% injected dose/g) at 24 h post-injection was significantly<br />

higher than the control (1.2G0.1%). A time-dependent biodistribution study showed sustained<br />

tumor accumulation of HPMA copolymer-RGD4C conjugate, efficient background clearance,<br />

and increasing T/B ratios over time. High T/B increases both tumor detection and therapeutic<br />

ratio. 178 Another advantage of using a polymeric conjugate of RGD was highlighted <strong>by</strong> comparing<br />

the biodistributions of HPMA copolymer-RGD4C conjugates and free RGD4C. 23 Organ distribution<br />

data in two murine xenograft human prostate models indicated higher tumor accumulation<br />

and lower extravasation in normal tissues for HPMA copolymer conjugates compared to free<br />

RGD4C peptides (Figure 9.4). Also, T/B was significantly higher for the macromolecular conjugate.<br />

These conjugates may be particularly advantageous for cancer radiotherapy because the<br />

combination of polyvalent interaction and EPR effect would help to retain the conjugate in the<br />

tumor, enhancing the radiation dose. 108,158<br />

The use of a water-soluble polymer (HPMA)-based conjugate of RGD peptide and the acyclic<br />

chelator cyclohexyl-diethylenetriamine pentaacetic acid (CHX-A 00 -DTPA) for angiogenesis<br />

directed ( 90 Y) radiotherapy has been studied. 108 After intravenous injection in prostate carcinoma<br />

(a) (b)<br />

FIGURE 9.3 (See color insert following page 522.) Scintigraphic images of human prostate tumor-bearing<br />

SCID mice 24 h post-intravenous injection of 99m Tc labeled HPMA copolymer conjugates. HPMA copolymer-<br />

RGD4C conjugate shows higher localization in tumor as compared to the control, HPMA copolymer-RGE4C<br />

conjugate (solid arrow). Additional higher activity was found in the kidney (broken arrow). The mouse<br />

radiograph (center) shows anatomic correlation of tumor and other organs. Figure legends: (a) HPMA copolymer-RGD4C<br />

conjugates; (b) HPMA copolymer-RGE4C conjugates. (From Mitra, A., et al., J. Control.<br />

Release, 102, 191, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


172<br />

% ID/g<br />

(a)<br />

% ID/g<br />

(b)<br />

16.000<br />

14.000<br />

12.000<br />

10.000<br />

8.000<br />

6.000<br />

4.000<br />

2.000<br />

0.000<br />

16.000<br />

14.000<br />

12.000<br />

10.000<br />

8.000<br />

6.000<br />

4.000<br />

2.000<br />

0.000<br />

HPMA-RGD4C (n=6)<br />

RGD4C-DPK (n=6)<br />

HPMA-RGE4C (n=6)<br />

RGE4C-DPK (n=6)<br />

Blood Heart Lung Liver Spleen Kidney Muscle Tumor<br />

HPMA-RGD4C (n=6)<br />

RGD4C-DPK (n=6)<br />

HPMA-RGE4C (n=6)<br />

RGE4C-DPK (n=6)<br />

Nanotechnology for Cancer Therapy<br />

Blood Heart Lung Liver Spleen Kidney Muscle Tumor<br />

FIGURE 9.4 Residual radioactivity in % injected dose per gram of organ tissue (%ID/g) 24 h post-intravenous<br />

injection of 99m Tc labeled HPMA copolymer conjugates and free peptides in (a) DU145 and (b) PC-3 prostate<br />

tumor xenograft-bearing SCID mice. The HPMA copolymers showed significantly higher tumor accumulation<br />

and reduced background localization than the peptides. The organ data are expressed as meanGSD (number of<br />

animals/group is shown). (From Line, B. R., et al., J. Nucl. Med., 46, 1152, 2005. With permission.)<br />

xenograft bearing SCID mice, the tumor accumulation of the conjugate peaked at 72 h post-injection,<br />

whereas the accumulation in other major organs significantly decreased during that period. A single<br />

injection of the 90 Y labeled conjugate at dose levels of 100 and 250 Ci caused significant reduction<br />

of tumor volume as compared to the untreated control that was evident from day 7 post-injection<br />

(Figure 9.5, top panel). Tumor histopathology showed cellular apoptosis in the treated groups,<br />

whereas no signs of acute toxicity were observed in the kidney, liver, and spleen (Figure 9.5, bottom<br />

panel).<br />

The RGD4C peptide has a doubly cyclized structure containing two disulphide bonds that<br />

affords high binding affinity and specificity for the aVb3 integrin. 178,179 However, the solution<br />

instability of the RGD4C disulphide bonds is a potential disadvantage that can lead to significant<br />

reduction in aVb3 binding affinity. 180 To overcome this problem, highly stable RGD peptides<br />

having monocyclic structure with head-to-tail cyclization containing a D-amino acid, e.g.,<br />

RGDfK or RGDyK, have been synthesized. 117 These have high affinity for aVb3 and have been<br />

used for delivery of imaging agents and therapeutics to tumor angiogenesis. 19,21 The long term (up<br />

to 192 h) biodistribution and tumor targeting properties of multivalent HPMA copolymer<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 173<br />

Tumor volume (cm 3 )<br />

2.00<br />

1.80<br />

1.60<br />

1.40<br />

1.20<br />

1.00<br />

0.80<br />

0.60<br />

0.40<br />

0.20<br />

0.00<br />

0 2 4 6 8 10 12 14 16 18 20 22<br />

Days post-treatment<br />

(a) (b)<br />

(c) (d)<br />

FIGURE 9.5 (See color insert following page 522.) TOP PANEL: Effect of 90 Y labeled HPMA copolymer-<br />

RGD4C conjugate treatment on human prostate tumor (DU145) growth in SCID mice. Animal groups treated<br />

with single dose of 100 mCi (p!0.03) and 250 mCi (p!0.01) 90 Y-HPMA-RGD4C conjugate showed significant<br />

tumor growth reduction as compared to the untreated controls <strong>by</strong> day seven post-treatment. Data are<br />

presented as mean of tumor volume (cm 3 )GSD (nZ6 mice per group). Figure legends: untreated control<br />

(closed square), 250 mCi of 90 Y-HPMA copolymer-RGD4C conjugate (closed circle), and 100 mCi of 90 Y-<br />

HPMA copolymer-RGD4C conjugate (open triangle). BOTTOM PANEL: Histological analysis of kidney and<br />

tumor samples at 21 days post-treatment following injection of 250 mCi 90 Y labeled HPMA copolymer-<br />

RGD4C conjugate or untreated control. Kidney samples taken from (a) control group and (b) 250 mCi treatment<br />

group were similar and showed no radiation induced toxicity. There was a complete lack of evidence of<br />

any tubular epithelial injury. Normal glomerular and proximal tubular anatomy was evident in both control and<br />

90 Y treated specimens. (c) The tumor sections from the control animals showed high grade epithelial malignancy<br />

typical of DU145 xenografts. (d) Tumor samples from 250 mCi treatment animals showed large cellular<br />

drop out areas (black arrow), higher numbers of eosinophilic cytoplasmic hyaline globular bodies (thanatosomes,<br />

open arrow), and pronounced nuclear atypia (hatched arrow) indicative of treatment effect/induced cell<br />

damage (From Mitra, A., et al., Nucl. Med. Biol., 33, 43, <strong>2006</strong>. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


174<br />

conjugates of RGD4C and RGDfK peptides have been studied. 181 Scintigraphic images and<br />

necropsy organ counts (Figure 9.6) showed that tumor accumulation of both HPMA-RGD4C<br />

and HPMA-RGDfK conjugates increased over time with peak accumulations at 4.9G0.9% (96 h<br />

p.i.) and 5.0G1.2% (48 h p.i.) ID/g, respectively. In contrast, the background organ distribution<br />

(a)<br />

(b)<br />

(c)<br />

% ID/g<br />

% ID/g<br />

10.000<br />

9.000<br />

8.000<br />

7.000<br />

6.000<br />

5.000<br />

4.000<br />

3.000<br />

2.000<br />

1.000<br />

0.000<br />

10.000<br />

9.000<br />

8.000<br />

7.000<br />

6.000<br />

5.000<br />

4.000<br />

3.000<br />

2.000<br />

1.000<br />

0.000<br />

15min (n=3)<br />

1h (n=3)<br />

6h (n=3)<br />

24h (n=4)<br />

48h (n=4)<br />

96h (n=4)<br />

192h (n=3)<br />

Blood Heart Lung Liver Spleen Kidney Small<br />

lntestine<br />

15min (n=3)<br />

1h (n=3)<br />

6h (n=3)<br />

24h (n=4)<br />

48h (n=4)<br />

96h (n=4)<br />

192h (n=3)<br />

Blood Heart Lung Liver Spleen Kidney Small<br />

lntestine<br />

Nanotechnology for Cancer Therapy<br />

Large<br />

lntestine<br />

Large<br />

lntestine<br />

Stomach Tail Muscle Tumor<br />

Stomach Tail Muscle Tumor<br />

FIGURE 9.6 (See color insert following page 522.) (a) Scintigraphic images of Lewis lung carcinomabearing<br />

mice up to 192 h post-intravenous injection of 111 In labeled copolymers. (1) HPMA copolymer-<br />

RGD4C conjugate and (2) HPMA copolymer-RGDfK conjugate showed marked localization in tumor at<br />

24 h p.i. and thereafter (solid arrow). The mouse radiograph shows anatomic correlation of tumor and other<br />

organs. Biodistribution of (b) 111 In-HPMA copolymer-RGD4C conjugate and (c) 111 In-HPMA copolymer-<br />

RGDfK conjugate. The organ activities, expressed as % injected dose per gram of organ tissue (%ID/g),<br />

showed persistent tumor localization and clearance of activity from the background organs (From Mitra,<br />

A., et al., J. Control. Release, 114, 175, <strong>2006</strong>. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 175<br />

rapidly cleared over time, resulting in significant increase in T/B ratios. The radioactive dose to<br />

organs as indicated <strong>by</strong> the area under curve was highest for the tumor. The polymer conjugates of<br />

RGD4C or RGDfK provides a means to enhance tumor uptake, decrease background accumulation,<br />

and enable selective delivery of therapeutic or diagnostic agents to tumor sites. Because both<br />

HPMA-RGD4C and HPMA-RGDfK conjugates have similar tumor targeting abilities and pharmacokinetics,<br />

RGDfK can be a suitable ligand because of higher solution stability and easier synthetic<br />

manipulation than the 12 amino-acid RGD4C.<br />

These studies with HPMA copolymer-RGD conjugates demonstrate the feasibility and advantages<br />

of using multivalent polymer-peptide conjugates. These conjugates show prolonged retention<br />

at the tumor site and enhanced T/B ratios. This increased contrast at the tumor site will be necessary<br />

for development of a clinically relevant diagnostic agent for therapy planning. Further, the<br />

improved therapeutic index will be ideal for angiogenesis directed chemo- or radiotherapy.<br />

9.5 SUMMARY AND FUTURE DIRECTIONS<br />

There is a tremendous interest in targeted drug delivery to tumor vasculature given the genetic<br />

stability and accessibility of angiogenesis markers and the importance of angiogenesis in tumor<br />

growth and metastasis. A number of ligands targeting angiogenic markers have been identified, but<br />

the major focus has been on targeting aVb3 integrins using RGD peptides. 21 Many reports of<br />

conjugating RGD ligands to a diverse group of macromolecular carrier systems have been<br />

published. These include water-soluble polymers (e.g., PEG, 139 N-(2-hydroxypropyl) methacrylamide,<br />

178 and polyethylenimine 164 ), proteins, 182,183 liposomes, 184,185 and nanoparticles. 165,186<br />

With respect to future development, a broad range of angiogenesis targeting polymer conjugates<br />

can be envisioned. As additional vascular targets are identified <strong>by</strong> in vitro and in vivo<br />

methods, angiogenesis targeting may become a critically important strategy for fighting neoplastic<br />

as well as non-neoplastic diseases. The first successful human trials for imaging aVb3 integrins<br />

using [ 18 F]-galacto-RGD 124,125 and the results of pre-clinical studies using macromolecular conjugates<br />

of RGD argue the validity of this concept.<br />

It is expected that the biokinetics of copolymer-conjugates will be continuously improved <strong>by</strong><br />

tailoring the molecular weight 187–190 and/or electronegative charge 190,191 of the conjugates. This<br />

flexibility in the design and synthesis of HPMA and other copolymer conjugates is likely to be<br />

important in modifying the blood half-life or reducing non-specific accumulation in normal organs,<br />

factors that should increase the conjugate therapeutic index. These may incorporate either enzymatically<br />

hydrolysable (e.g., tetrapeptide GFLG) 143 or pH sensitive (e.g., hydrazone linker) 192<br />

linkers for intracellular drug release. In addition, they will likely demonstrate targeting ligand<br />

multivalency to increase the tumor uptake and therapeutic efficacy.<br />

It is also expected that there will be expanded use of therapeutic radionuclides to destroy<br />

angiogenic vasculature and surrounding tumor cells. Because radioactive emissions can kill at a<br />

distance from the point of radioisotope localization, they have a diameter of effectiveness that may<br />

overcome the problem of tumor heterogeneity that has plagued other molecular therapies. In short,<br />

the ever increasing interest in polymer-based therapeutics promises continued progress in angiogenesis<br />

targeted, tumor imaging, and therapy.<br />

9.6 LIST OF ABBREVIATIONS<br />

AFK Ala-Phe-Lys<br />

APA Aminopeptidase A<br />

APMA-CHX-A 00 -DTPA N-methacryloylaminopropyl-2-amino-3-(isothiourea-phenyl)propyl-cyclohexane-1,2-diamine-N,N-N<br />

0 ,N 0 ,N 00 ,N 00 -<br />

pentaacetic acid<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


176<br />

Nanotechnology for Cancer Therapy<br />

APN Aminopeptidase N<br />

BBB Blood–brain barrier<br />

CAT Chloramphenicol acetyl transferase<br />

CDTRL Cys-Asp-Thr-Arg-Leu<br />

CGKRK Cys-Gly-Lys-Arg-Lys<br />

CHX-A 00 -DTPA Cyclohexyl–diethylenetriamine pentaacetic acid<br />

CPRECESIC Cys-Pro-Arg-Glu-Cys-Glu-Ser-Ile-Cys<br />

CRGRRST Cys-Arg-Gly-Arg-Arg-Ser-Thr<br />

CSRPRRSEC Cys-Ser-Arg-Pro-Arg-Arg-Ser-Glu-Cys<br />

DOTA 1,4,7,10-tetra-azacylcododecane-N,N 0 ,N 00 ,N 000 -tetraacetic acid<br />

DOX Doxorubicin<br />

DPK N-u-bis(2-pyridylmethyl)-L-lysine<br />

DTPA Diethylenetriaminepentaacetic acid<br />

ECM Extracellular matrix<br />

EPR Enhanced permeability and retention<br />

FB Fluorobenzoyl<br />

FDA Food and drug administration<br />

FDG 2-Fluoro-2-deoxyglucose<br />

5-Fu 5-Fluorouracil<br />

GFLG Gly-Phe-Leu-Gly<br />

GPLGV Gly-Pro-Leu-Gly-Val<br />

GPLGIAGQ Gly-Pro-Leu-Gly-Ile-Ala-Gly-Gln<br />

GPLGVRGK Gly-Pro-Leu-Gly-Val-Arg-Gly-Lys<br />

GVPLSLTMGC Gly-Val-Pro-Leu-Ser-Leu-Thr-Met-Gly-Cys<br />

HPMA N-(2-hydroxypropyl) methacrylamide<br />

HWGF His-Trp-Gly-Phe<br />

HYNIC Hydrazinonicotinamide<br />

MA-GG-DPK N-methacryloylglycylglycyl-(N-u-bis(2-pyridylmethyl)-L-lysine)<br />

MA-GG-RGD4C N-methacryloylglycylglycyl-RGD4C<br />

MA-Tyr N-methacryloyltyrosinamide<br />

MBq Megabecquerel<br />

mCi Millicurie<br />

MMP Matrix metalloproteinases<br />

MRI Magnetic resonance imaging<br />

NGR Asn-Gly-Arg<br />

NIR Near-infrared<br />

PEG Polyethylene glycol<br />

PEI Polyethylenimine<br />

PET Positron emission tomography<br />

PLGL Pro-Leu-Gly-Leu<br />

PLL Poly-L-lysine<br />

RGD Arg-Gly-Asp<br />

RGE Arg-Gly-Glu<br />

SCID Severe combined immunodeficient<br />

SMSIARL Ser-Met-Ser-Ile-Ala-Arg-Leu<br />

T/B Tumor-to-background<br />

TIMP Tissue inhibitors of matrix metalloproteinases<br />

VEGF Vascular endothelial growth factor<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 177<br />

ACKNOWLEDGMENT<br />

Support from the American Russian Cancer Alliance and National Institutes of Health (1 R01<br />

EB0207171) is acknowledged.<br />

REFERENCES<br />

1. Ferrara, N. and Kerbel, R. S., Angiogenesis as a therapeutic target, Nature, 438, 967, 2005.<br />

2. Carmeliet, P., Angiogenesis in health and disease, Nat. Med., 9, 653, 2003.<br />

3. Carmeliet, P. and Jain, R. K., Angiogenesis in cancer and other diseases, Nature, 407, 249, 2000.<br />

4. Carmeliet, P., Mechanisms of angiogenesis and arteriogenesis, Nat. Med., 6, 389, 2000.<br />

5. Thorpe, P. E., Vascular targeting agents as cancer therapeutics, Clin. Cancer Res., 10, 415, 2004.<br />

6. Siemann, D. W. et al., Differentiation and definition of vascular-targeted therapies, Clin. Cancer<br />

Res., 11, 416, 2005.<br />

7. Kerbel, R. and Folkman, J., Clinical translation of angiogenesis inhibitors, Nat. Rev. Cancer, 2, 727,<br />

2002.<br />

8. Ellis, L. M., Synopsis of angiogenesis inhibitors in oncology, Oncology (Huntingt), 16, 14, 2002.<br />

9. Risau, W., Mechanisms of angiogenesis, Nature, 386, 671, 1997.<br />

10. Gasparini, G. et al., Angiogenic inhibitors: A new therapeutic strategy in oncology, Nat. Clin. Pract.<br />

Oncol., 2, 562, 2005.<br />

11. Folkman, J., Angiogenesis inhibitors: A new class of drugs, Cancer Biol. Ther., 2, S127, 2003.<br />

12. Ranieri, G. and Gasparini, G., Angiogenesis and angiogenesis inhibitors: A new potential anticancer<br />

therapeutic strategy, Curr. Drug Targets Immune Endocr. Metabol. Disord., 1, 241, 2001.<br />

13. Ferrara, N. et al., Discovery and development of bevacizumab, an anti-VEGF antibody for treating<br />

cancer, Nat. Rev. Drug Discov., 3, 391, 2004.<br />

14. Tozer, G. M., Kanthou, C., and Baguley, B. C., Disrupting tumour blood vessels, Nat. Rev. Cancer,<br />

5, 423, 2005.<br />

15. Narazaki, M. and Tosato, G., Targeting coagulation to the tumor microvasculature: Perspectives and<br />

therapeutic implications from preclinical studies, J. Natl. Cancer Inst., 97, 705, 2005.<br />

16. Ching, L. M. et al., Induction of intratumoral tumor necrosis factor (TNF) synthesis and hemorrhagic<br />

necrosis <strong>by</strong> 5,6-dimethylxanthenone-4-acetic acid (DMXAA) in TNF knockout mice, Cancer Res.,<br />

59, 3304, 1999.<br />

17. Tozer, G. M. et al., The biology of the combretastatins as tumour vascular targeting agents, Int.<br />

J. Exp. Pathol., 83, 21, 2002.<br />

18. Pedley, R. B. et al., Eradication of colorectal xenografts <strong>by</strong> combined radioimmunotherapy and<br />

combretastatin a-4 3-O-phosphate, Cancer Res., 61, 4716, 2001.<br />

19. Haubner, R. and Wester, H. J., Radiolabeled tracers for imaging of tumor angiogenesis and evaluation<br />

of anti-angiogenic therapies, Curr. Pharm. Des., 10, 1439, 2004.<br />

20. Li, W. P. and Anderson, C. J., Imaging matrix metalloproteinase expression in tumors, Q. J. Nucl.<br />

Med., 47, 201, 2003.<br />

21. Temming, K. et al., RGD-based strategies for selective delivery of therapeutics and imaging agents<br />

to the tumour vasculature, Drug Resist.Updat., 8 (6), 381–402, 2005.<br />

22. Maeda, H. et al., Tumor vascular permeability and the EPR effect in macromolecular therapeutics: A<br />

review, J. Control. Release, 65, 271, 2000.<br />

23. Line, B. R. et al., Targeting tumor angiogenesis: Comparison of peptide and polymer-peptide<br />

conjugates, J. Nucl. Med., 46, 1552, 2005.<br />

24. Mahadevan, V. and Hart, I. R., Metastasis and angiogenesis, Acta Oncol., 29, 97, 1990.<br />

25. Hanahan, D. and Folkman, J., Patterns and emerging mechanisms of the angiogenic switch during<br />

tumorigenesis, Cell, 86, 353, 1996.<br />

26. Folkman, J., Tumor angiogenesis: Therapeutic implications, N. Engl. J. Med., 285, 1182, 1971.<br />

27. Folkman, J., What is the evidence that tumors are angiogenesis dependent?, J. Natl. Cancer Inst., 82,<br />

4, 1990.<br />

28. Weidner, N., Tumoural vascularity as a prognostic factor in cancer patients: The evidence continues<br />

to grow, J. Pathol., 184, 119, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


178<br />

Nanotechnology for Cancer Therapy<br />

29. Weidner, N., Tumour vascularity and proliferation: Clear evidence of a close relationship, J. Pathol.,<br />

189, 297, 1999.<br />

30. Bergers, G. and Benjamin, L. E., Tumorigenesis and the angiogenic switch, Nat. Rev. Cancer, 3, 401,<br />

2003.<br />

31. Hanahan, D. and Weinberg, R. A., The hallmarks of cancer, Cell, 100, 57, 2000.<br />

32. Folkman, J., Endogenous angiogenesis inhibitors, APMIS, 112, 496, 2004.<br />

33. Nyberg, P., Xie, L., and Kalluri, R., Endogenous inhibitors of angiogenesis, Cancer Res., 65, 3967,<br />

2005.<br />

34. Sund, M. et al., Function of endogenous inhibitors of angiogenesis as endothelium-specific tumor<br />

suppressors, Proc. Natl. Acad. Sci. U.S.A., 102, 2934, 2005.<br />

35. Relf, M. et al., Expression of the angiogenic factors vascular endothelial cell growth factor, acidic<br />

and basic fibroblast growth factor, tumor growth factor beta-1, platelet-derived endothelial cell<br />

growth factor, placenta growth factor, and pleiotrophin in human primary breast cancer and its<br />

relation to angiogenesis, Cancer Res., 57, 963, 1997.<br />

36. Yancopoulos, G. D. et al., Vascular-specific growth factors and blood vessel formation, Nature, 407,<br />

242, 2000.<br />

37. Hanahan, D. et al., Transgenic mouse models of tumour angiogenesis: The angiogenic switch, its<br />

molecular controls, and prospects for preclinical therapeutic models, Eur. J Cancer, 32A, 2386,<br />

1996.<br />

38. Harris, A. L., Hypoxia—a key regulatory factor in tumour growth, Nat. Rev. Cancer, 2, 38, 2002.<br />

39. Semenza, G. L., HIF-1 and tumor progression: Pathophysiology and therapeutics, Trends Mol. Med.,<br />

8, S62, 2002.<br />

40. Jain, R. K., Munn, L. L., and Fukumura, D., Dissecting tumour pathophysiology using intravital<br />

microscopy, Nat. Rev. Cancer, 2, 266, 2002.<br />

41. Morikawa, S. et al., Abnormalities in pericytes on blood vessels and endothelial sprouts in tumors,<br />

Am. J. Pathol., 160, 985, 2002.<br />

42. Jain, R. K., Molecular regulation of vessel maturation, Nat. Med., 9, 685, 2003.<br />

43. Baluk, P., Hashizume, H., and McDonald, D. M., Cellular abnormalities of blood vessels as targets in<br />

cancer, Curr. Opin. Genet. Dev., 15, 102, 2005.<br />

44. Ruoslahti, E., Specialization of tumour vasculature, Nat. Rev. Cancer, 2, 83, 2002.<br />

45. Brekken, R. A. and Thorpe, P. E., Vascular endothelial growth factor and vascular targeting of solid<br />

tumors, Anticancer Res., 21, 4221, 2001.<br />

46. Burrows, F. J. and Thorpe, P. E., Vascular targeting—a new approach to the therapy of solid tumors,<br />

Pharmacol. Ther., 64, 155, 1994.<br />

47. Neri, D. and Bicknell, R., Tumour vascular targeting, Nat. Rev. Cancer, 5, 436, 2005.<br />

48. Pasqualini, R. and Ruoslahti, E., Organ targeting in vivo using phage display peptide libraries,<br />

Nature, 380, 364, 1996.<br />

49. Arap, W. et al., Steps toward mapping the human vasculature <strong>by</strong> phage display, Nat. Med., 8, 121,<br />

2002.<br />

50. Ruoslahti, E., Vascular zip codes in angiogenesis and metastasis, Biochem. Soc. Trans., 32, 397,<br />

2004.<br />

51. Haubner, R. H. et al., Radiotracer-based strategies to image angiogenesis, Q. J. Nucl. Med., 47, 189,<br />

2003.<br />

52. McQuade, P. and Knight, L. C., Radiopharmaceuticals for targeting the angiogenesis marker<br />

alpha(v)beta(3), Q. J. Nucl. Med., 47, 209, 2003.<br />

53. Pasqualini, R., Koivunen, E., and Ruoslahti, E., Alpha v integrins as receptors for tumor targeting <strong>by</strong><br />

circulating ligands, Nat. Biotechnol., 15, 542, 1997.<br />

54. Pasqualini, R. et al., Aminopeptidase N is a receptor for tumor-homing peptides and a target for<br />

inhibiting angiogenesis, Cancer Res., 60, 722, 2000.<br />

55. Isberg, R. R. and Tran, V. N., Binding and internalization of microorganisms <strong>by</strong> integrin receptors,<br />

Trends Microbiol., 2, 10, 1994.<br />

56. Hart, S. L. et al., Cell binding and internalization <strong>by</strong> filamentous phage displaying a cyclic Arg-Gly-<br />

Asp-containing peptide, J. Biol. Chem., 269, 12468, 1994.<br />

57. Ferrara, N., Gerber, H. P., and LeCouter, J., The biology of VEGF and its receptors, Nat. Med., 9,<br />

669, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 179<br />

58. Ferrara, N., Vascular endothelial growth factor: Basic science and clinical progress, Endocr. Rev.,<br />

25, 581, 2004.<br />

59. Ferrara, N., Vascular endothelial growth factor as a target for anticancer therapy, Oncologist, 9<br />

(Suppl 1), 2, 2004.<br />

60. Ahmed, S. I., Thomas, A. L., and Steward, W. P., Vascular endothelial growth factor (VEGF)<br />

inhibition <strong>by</strong> small molecules, J. Chemother., 16 (Suppl 4), 59, 2004.<br />

61. Sato, Y., Aminopeptidases and angiogenesis, Endothelium, 10, 287, 2003.<br />

62. Chang, Y. W. et al., CD13 (aminopeptidase N) can associate with tumor-associated antigen L6 and<br />

enhance the motility of human lung cancer cells, Int. J. Cancer, 116, 243, 2005.<br />

63. Ishii, K. et al., Aminopeptidase N regulated <strong>by</strong> zinc in human prostate participates in tumor cell<br />

invasion, Int. J. Cancer, 92, 49, 2001.<br />

64. Arap, W., Pasqualini, R., and Ruoslahti, E., Cancer treatment <strong>by</strong> targeted drug delivery to tumor<br />

vasculature in a mouse model, Science, 279, 377, 1998.<br />

65. Eller<strong>by</strong>, H. M. et al., Anti-cancer activity of targeted pro-apoptotic peptides, Nat. Med., 5, 1032,<br />

1999.<br />

66. Schlingemann, R. O. et al., Aminopeptidase a is a constituent of activated pericytes in angiogenesis,<br />

J. Pathol., 179, 436, 1996.<br />

67. Marchio, S. et al., Aminopeptidase A is a functional target in angiogenic blood vessels, Cancer Cell,<br />

5, 151, 2004.<br />

68. Vihinen, P. and Kahari, V. M., Matrix metalloproteinases in cancer: Prognostic markers and therapeutic<br />

targets, Int. J. Cancer, 99, 157, 2002.<br />

69. Vihinen, P., Ala-aho, R., and Kahari, V. M., Matrix metalloproteinases as therapeutic targets in<br />

cancer, Curr. Cancer Drug Targets, 5, 203, 2005.<br />

70. Egeblad, M. and Werb, Z., New functions for the matrix metalloproteinases in cancer progression,<br />

Nat. Rev. Cancer, 2, 161, 2002.<br />

71. Nguyen, M., Arkell, J., and Jackson, C. J., Human endothelial gelatinases and angiogenesis, Int.<br />

J. Biochem. Cell Biol., 33, 960, 2001.<br />

72. Himelstein, B. P. et al., Metalloproteinases in tumor progression: The contribution of MMP-9, Invas.<br />

Metast., 14, 246, 1994.<br />

73. Koivunen, E. et al., Tumor targeting with a selective gelatinase inhibitor, Nat. Biotechnol., 17, 768,<br />

1999.<br />

74. Hidalgo, M. and Eckhardt, S. G., Development of matrix metalloproteinase inhibitors in cancer<br />

therapy, J. Natl. Cancer Inst., 93, 178, 2001.<br />

75. Levitt, N. C. et al., Phase I and pharmacological study of the oral matrix metalloproteinase inhibitor,<br />

MMI270 (CGS27023A), in patients with advanced solid cancer, Clin. Cancer Res., 7, 1912, 2001.<br />

76. Eatock, M. et al., A dose-finding and pharmacokinetic study of the matrix metalloproteinase<br />

inhibitor MMI270 (previously termed CGS27023A) with 5-FU and folinic acid, Cancer Chemother.<br />

Pharmacol., 55, 39, 2005.<br />

77. Kline, T., Torgov, M. Y., Mendelsohn, B. A., Cerveny, C. G., and Senter, P. D., Novel antitumor<br />

prodrugs designed for activation <strong>by</strong> matrix metalloproteinases-2 and -9, Mol. Pharm., 1, 9, 2004.<br />

78. Albright, C. F. et al., Matrix metalloproteinase-activated doxorubicin prodrugs inhibit HT1080<br />

xenograft growth better than doxorubicin with less toxicity, Mol. Cancer Ther., 4, 751, 2005.<br />

79. Bae, M. et al., Metalloprotease-specific poly(ethylene glycol) methyl ether–peptide–doxorubicin<br />

conjugate for targeting anticancer drug delivery based on angiogenesis, Drugs Exp. Clin. Res., 29,<br />

15, 2003.<br />

80. Kratz, F. et al., Development and in vitro efficacy of novel MMP2 and MMP9 specific doxorubicin<br />

albumin conjugates, Bioorg. Med. Chem. Lett., 11, 2001, 2001.<br />

81. Kratz, F. et al., Probing the cysteine-34 position of endogenous serum albumin with thiol-binding<br />

doxorubicin derivatives. Improved efficacy of an acid-sensitive doxorubicin derivative with specific<br />

albumin-binding properties compared to that of the parent compound, J. Med. Chem., 45, 5523,<br />

2002.<br />

82. Mansour, A. M. et al., A new approach for the treatment of malignant melanoma: Enhanced antitumor<br />

efficacy of an albumin-binding doxorubicin prodrug that is cleaved <strong>by</strong> matrix<br />

metalloproteinase 2, Cancer Res., 63, 4062, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


180<br />

Nanotechnology for Cancer Therapy<br />

83. Kuhnast, B. et al., Targeting of gelatinase activity with a radiolabeled cyclic HWGF peptide, Nucl.<br />

Med. Biol., 31, 337, 2004.<br />

84. Zheng, Q. H. et al., Synthesis and preliminary biological evaluation of MMP inhibitor radiotracers<br />

[ 11 C]methyl-halo-CGS 27023A analogs, new potential PET breast cancer imaging agents, Nucl.<br />

Med. Biol., 29, 761, 2002.<br />

85. Furumoto, S., Iwata, R., and Ido, T., Design and synthesis of fluorine-18 labeled matrix metalloproteinase<br />

inhibitors for cancer imaging, J. Labelled Compd. Rad., 45, 975, 2002.<br />

86. Fei, X., Hutchins, G. D. et al., Synthesis of radiolabeled biphenylsulfonamide matrix metalloproteinase<br />

inhibitors as new potential PET cancer imaging agents, Bioorg. Med. Chem. Lett., 13, 2217,<br />

2003.<br />

87. Furumoto, S. et al., Development of a new 18 F-labeled matrix metalloproteinase-2 inhibitor for<br />

cancer imaging <strong>by</strong> PET, J. Nucl. Med., 43 (Suppl), 364, 2002.<br />

88. Furumoto, S. et al., Tumor detection using 18 F-labeled matrix metalloproteinase-2 inhibitor, Nucl.<br />

Med. Biol., 30, 119, 2003.<br />

89. Zheng, Q. H. et al., Synthesis, biodistribution and micro-PET imaging of a potential cancer biomarker<br />

carbon-11 labeled MMP inhibitor ( 2 R)-2-[[4-(6-fluorohex-1-ynyl)phenyl]sulfonylamino]-3methylbutyric<br />

acid [ 11 C]methyl ester, Nucl. Med. Biol., 30, 753, 2003.<br />

90. Li, W. P. et al., 64 Cu-DOTA-CTTHWGFTLC: A selective gelatinase inhibitor for tumor imaging,<br />

J. Nucl. Med., 43 (Suppl), 227, 2002.<br />

91. Li, W. P. et al., In vitro and in vivo evaluation of a radiolabeled gelatinase inhibitor for microPET<br />

imaging of metastatic breast cancer, Mol. Imaging Biol., 4 (Suppl), 23, 2002.<br />

92. Giersing, B. K. et al., Synthesis and characterization of 111 In-DTPA-N-TIMP-2: A radiopharmaceutical<br />

for imaging matrix metalloproteinase expression, Bioconjug. Chem., 12, 964, 2001.<br />

93. Kulasegaram, R. et al., In vivo evaluation of 111 In-DTPA-N-TIMP-2 in Kaposi sarcoma associated<br />

with HIV infection, Eur. J. Nucl. Med., 28, 756, 2001.<br />

94. Bremer, C., Tung, C. H., and Weissleder, R., In vivo molecular target assessment of matrix metalloproteinase<br />

inhibition, Nat. Med., 7, 743, 2001.<br />

95. Bremer, C. et al., Optical imaging of matrix metalloproteinase-2 activity in tumors: Feasibility study<br />

in a mouse model, Radiology, 221, 523, 2001.<br />

96. Bremer, C., Tung, C. H., and Weissleder, R., Molecular imaging of MMP expression and therapeutic<br />

MMP inhibition, Acad. Radiol., 9 (Suppl 2), S314, 2002.<br />

97. Pham, W. et al., Developing a peptide-based near-infrared molecular probe for protease sensing,<br />

Bioconjug. Chem., 15, 1403, 2004.<br />

98. Cox, D. et al., The pharmacology of the integrins, Med. Res. Rev., 14, 195, 1994.<br />

99. van Der, F. A. and Sonnenberg, A., Function and interactions of integrins, Cell Tissue Res., 305, 285,<br />

2001.<br />

100. Sepulveda, J. L., Gkretsi, V., and Wu, C., Assembly and signaling of adhesion complexes, Curr. Top.<br />

Dev. Biol., 68, 183, 2005.<br />

101. Hood, J. D. and Cheresh, D. A., Role of integrins in cell invasion and migration, Nat. Rev. Cancer,2,<br />

91, 2002.<br />

102. Giancotti, F. G. and Ruoslahti, E., Integrin signaling, Science, 285, 1028, 1999.<br />

103. Brooks, P. C., Clark, R. A., and Cheresh, D. A., Requirement of vascular integrin alpha v beta 3 for<br />

angiogenesis, Science, 264, 569, 1994.<br />

104. Friedlander, M. et al., Definition of two angiogenic pathways <strong>by</strong> distinct alpha v integrins, Science,<br />

270, 1500, 1995.<br />

105. Eliceiri, B. P. and Cheresh, D. A., Role of alpha v integrins during angiogenesis, Cancer J., 6 (Suppl<br />

3), S245, 2000.<br />

106. Stupack, D. G. and Cheresh, D. A., Integrins and angiogenesis, Curr. Top. Dev. Biol., 64, 207, 2004.<br />

107. Janssen, M. L. et al., Tumor targeting with radiolabeled alpha(v)beta(3) integrin binding peptides in a<br />

nude mouse model, Cancer Res., 62, 6146, 2002.<br />

108. Mitra, A. et al., Polymer-peptide conjugates for angiogenesis targeted tumor radiotherapy, Nucl.<br />

Med. Biol., 33, 43, <strong>2006</strong>.<br />

109. Goodman, S. L. et al., Nanomolar small molecule inhibitors for alphav(beta)6, alphav(beta)5, and<br />

alphav(beta)3 integrins, J. Med. Chem., 45, 1045, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 181<br />

110. Winter, P. M. et al., Molecular imaging of angiogenesis in nascent Vx-2 rabbit tumors using a novel<br />

alpha(nu)beta3-targeted nanoparticle and 1.5 tesla magnetic resonance imaging, Cancer Res., 63,<br />

5838, 2003.<br />

111. Chen, X., Conti, P. S., and Moats, R. A., In vivo near-infrared fluorescence imaging of integrin<br />

alphavbeta3 in brain tumor xenografts, Cancer Res., 64, 8009, 2004.<br />

112. Haubner, R., Finsinger, D., and Kessler, H., Stereoisomeric peptide libraries and peptidomimetics for<br />

designing selective inhibitors of the alphav beta3 integrin for a new cancer therapy, Angew. Chem.<br />

Int. Ed. Engl., 36, 1374, 1997.<br />

113. Aumailley, M. et al., Arg-Gly-Asp constrained within cyclic pentapeptides. Strong and selective<br />

inhibitors of cell adhesion to vitronectin and laminin fragment P1, FEBS Lett., 291, 50, 1991.<br />

114. Gurrath, M. et al., Conformation/activity studies of rationally designed potent anti-adhesive RGD<br />

peptides, Eur. J. Biochem., 210, 911, 1992.<br />

115. Haubner, R. et al., Structural and functional aspects of RGD containing cyclic pentapeptides as<br />

highly potent and selective avb3 antagonists, J. Am. Chem. Soc., 118, 7461, 1996.<br />

116. Haubner, R. et al., Comparison of tumor uptake and biokinetics of I-125 and F-18 labeled RGDpeptides,<br />

J. Labelled Compd. Rad., 42, S36, 1999.<br />

117. Haubner, R. et al., Radiolabeled alpha(v)beta3 integrin antagonists: A new class of tracers for tumor<br />

targeting, J. Nucl. Med., 40, 1061, 1999.<br />

118. Chen, X. et al., MicroPET and autoradiographic imaging of breast cancer alpha v-integrin expression<br />

using 18 F- and 64 Cu-labeled RGD peptide, Bioconjug. Chem., 15, 41, 2004.<br />

119. Wang, W. et al., Convenient solid-phase synthesis of diethylenetriaminepenta–acetic acid (DTPA)conjugated<br />

cyclic RGD peptide analogues, Cancer Biother. Radiopharm., 20, 547, 2005.<br />

120. Haubner, R. et al., Glycosylated RGD-containing peptides: Tracer for tumor targeting and angiogenesis<br />

imaging with improved biokinetics, J. Nucl. Med., 42, 326, 2001.<br />

121. Haubner, R. et al., Noninvasive imaging of alpha(v)beta3 integrin expression using 18 F-labeled<br />

RGD-containing glycopeptide and positron emission tomography, Cancer Res., 61, 1781, 2001.<br />

122. Haubner, R. et al., [ 18 F]Galacto-RGD: Synthesis, radiolabeling, metabolic stability, and radiation<br />

dose estimates, Bioconjug. Chem., 15, 61, 2004.<br />

123. Haubner, R. et al., [F-18]-RGD-peptides conjugated with hydrophilic tetrapeptides for the noninvasive<br />

determination of the avb3 integrin, J. Nucl. Med., 43 (Suppl), 89P, 2002.<br />

124. Haubner, R. et al., Noninvasive visualization of the activated alphavbeta3 integrin in cancer patients<br />

<strong>by</strong> positron emission tomography and [(18)F]Galacto-RGD, PLoS Med., 2, e70, 2005.<br />

125. Beer, A. J. et al., Biodistribution and pharmacokinetics of the alphavbeta3-selective tracer 18 Fgalacto-RGD<br />

in cancer patients, J. Nucl. Med., 46, 1333, 2005.<br />

126. Sipkins, D. A. et al., Detection of tumor angiogenesis in vivo <strong>by</strong> alphaVbeta3-targeted magnetic<br />

resonance imaging, Nat. Med., 4, 623, 1998.<br />

127. Anderson, S. A. et al., Magnetic resonance contrast enhancement of neovasculature with alpha(v)beta(3)-targeted<br />

nanoparticles, Magn. Reson. Med., 44, 433, 2000.<br />

128. Winter, P. M. et al., Molecular imaging of angiogenesis in early-stage atherosclerosis with alpha(v)beta3-integrin-targeted<br />

nanoparticles, Circulation, 108, 2270, 2003.<br />

129. Cheng, Z. et al., Near-infrared fluorescent RGD peptides for optical imaging of integrin alphavbeta3<br />

expression in living mice, Bioconjug. Chem., 16, 1433, 2005.<br />

130. de Groot, F. M. et al., Design, synthesis, and biological evaluation of a dual tumor-specific motive<br />

containing integrin-targeted plasmin-cleavable doxorubicin prodrug, Mol. Cancer Ther., 1, 901,<br />

2002.<br />

131. Devy, L. et al., Plasmin-activated doxorubicin prodrugs containing a spacer reduce tumor growth<br />

and angiogenesis without systemic toxicity, FASEB J., 18, 565, 2004.<br />

132. Kim, J. W. and Lee, H. S., Tumor targeting <strong>by</strong> doxorubicin-RGD-4C peptide conjugate in an<br />

orthotopic mouse hepatoma model, Int. J. Mol. Med., 14, 529, 2004.<br />

133. Janssen, M. et al., Comparison of a monomeric and dimeric radiolabeled RGD-peptide for tumor<br />

targeting, Cancer Biother. Radiopharm., 17, 641, 2002.<br />

134. Liu, S. et al., 99m Tc-labeling of a hydrazinonicotinamide-conjugated vitronectin receptor antagonist<br />

useful for imaging tumors, Bioconjug. Chem., 12, 624, 2001.<br />

135. Chen, X. et al., MicroPET imaging of breast cancer alphav-integrin expression with 64 Cu-labeled<br />

dimeric RGD peptides, Mol. Imaging Biol., 6, 350, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


182<br />

Nanotechnology for Cancer Therapy<br />

136. Chen, X. et al., Micro-PET imaging of alphavbeta3-integrin expression with 18 F-labeled dimeric<br />

RGD peptide, Mol. Imaging, 3, 96, 2004.<br />

137. Wu, Y. et al., MicroPET imaging of glioma integrin {alpha}v{beta}3 expression using (64)Culabeled<br />

tetrameric RGD peptide, J. Nucl. Med., 46, 1707, 2005.<br />

138. Poethko, T. et al., Two-step methodology for high-yield routine radiohalogenation of peptides:<br />

(18)F-labeled RGD and octreotide analogs, J. Nucl. Med., 45, 892, 2004.<br />

139. Chen, X. et al., Pharmacokinetics and tumor retention of 125 I-labeled RGD peptide are improved <strong>by</strong><br />

PEGylation, Nucl. Med. Biol., 31, 11, 2004.<br />

140. Chen, X. et al., MicroPET imaging of brain tumor angiogenesis with 18 F-labeled PEGylated RGD<br />

peptide, Eur. J. Nucl. Med. Mol. Imaging, 31, 1081, 2004.<br />

141. Chen, X. et al., Pegylated Arg-Gly-Asp peptide: 64 Cu labeling and PET imaging of brain tumor<br />

alphavbeta3-integrin expression, J. Nucl. Med., 45, 1776, 2004.<br />

142. Poethko, T. et al., Improved tumor uptake, tumor retention and tumor/background ratios of pegylated<br />

RGD-multimers, J. Nucl. Med., 44, 46P, 2003.<br />

143. Putnam, D. and Kopecek, J., Polymer conjugates with anticancer activity, Adv. Polym. Sci., 122, 56,<br />

1995.<br />

144. Kopecek, J. et al., HPMA copolymer-anticancer drug conjugates: Design, activity, and mechanism<br />

of action, Eur. J. Pharm. Biopharm., 50, 61, 2000.<br />

145. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2, 347, 2003.<br />

146. Greenwald, R. B. et al., Effective drug delivery <strong>by</strong> PEGylated drug conjugates, Adv. Drug Deliv.<br />

Rev., 55, 217, 2003.<br />

147. Kopecek, J. et al., Water soluble polymers in tumor targeted delivery, J. Control. Release, 74, 147,<br />

2001.<br />

148. Lu, Z. R. et al., Design of novel bioconjugates for targeted drug delivery, J. Control. Release, 78,<br />

165, 2002.<br />

149. Duncan, R. et al., Fate of N-(2-hydroxypropyl)methacrylamide copolymers with pendent galactosamine<br />

residues after intravenous administration to rats, Biochim. Biophys. Acta, 880, 62, 1986.<br />

150. Kopecek, J. and Duncan, R., Poly[N-(2-Hydroxypropyl)-methacrylamide] macromolecules as drug<br />

carrier systems, In Polymers in Controlled Drug Delivery, Illum, L. and Davis, S. S. Eds., Butterworth-Heinemann,<br />

Oxford, UK, 1987.<br />

151. Lu, Z. R. et al., Functionalized semitelechelic poly[N-(2-hydroxypropyl)methacrylamide] for protein<br />

modification, Bioconjug. Chem., 9, 793, 1998.<br />

152. Mitra, A., et al., Nanocarriers for nuclear imaging and radiotherapy of cancer, Curr. Pharm. Des.,in<br />

press, <strong>2006</strong>.<br />

153. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent<br />

smancs, Cancer Res., 46, 6387, 1986.<br />

154. Satchi-Fainaro, R. et al., Targeting angiogenesis with a conjugate of HPMA copolymer and TNP-<br />

470, Nat. Med., 10, 255, 2004.<br />

155. Satchi-Fainaro, R. et al., Inhibition of vessel permeability <strong>by</strong> TNP-470 and its polymer conjugate,<br />

caplostatin, Cancer Cell, 7, 251, 2005.<br />

156. Shiah, J. J. et al., Biodistribution of free and N-(2-hydroxypropyl)methacrylamide copolymer-bound<br />

mesochlorin e(6) and adriamycin in nude mice bearing human ovarian carcinoma OVCAR-3 xenografts,<br />

J. Control. Release, 61, 145, 1999.<br />

157. Mammen, M. et al., Polyvalent interactions in biological systems: Implications for design and use of<br />

multivalent ligands and inhibitors, Angew. Chem. Int. Ed., 37, 2754, 1998.<br />

158. Wester, H. J. and Kessler, H., Molecular targeting with peptides or Peptide-polymer conjugates: Just<br />

a question of size?, J. Nucl. Med., 46, 1940, 2005.<br />

159. Bhargava, P. et al., A Phase I and pharmacokinetic study of TNP-470 administered weekly to<br />

patients with advanced cancer, Clin. Cancer Res., 5, 1989, 1999.<br />

160. Dickson, P. V., Nathwani, A. C., and Davidoff, A. M., Delivery of antiangiogenic agents for cancer<br />

gene therapy, Technol. Cancer Res. Treat., 4, 331, 2005.<br />

161. Tandle, A., Blazer, D. G. III, and Libutti, S. K., Antiangiogenic gene therapy of cancer: Recent<br />

developments, J. Transl. Med., 2, 22, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Conjugates for Angiogenesis-Targeted Tumor Imaging and Therapy 183<br />

162. Witlox, A. M. et al., Conditionally replicative adenovirus with tropism expanded towards integrins<br />

inhibits osteosarcoma tumor growth in vitro and in vivo, Clin. Cancer Res., 10, 61, 2004.<br />

163. Okada, Y. et al., Optimization of antitumor efficacy and safety of in vivo cytokine gene therapy using<br />

RGD fiber-mutant adenovirus vector for preexisting murine melanoma, Biochim. Biophys. Acta,<br />

1670, 172, 2004.<br />

164. Kim, W. J. et al., Soluble Flt-1 gene delivery using PEI-g-PEG-RGD conjugate for anti-angiogenesis,<br />

J. Control. Release, 106, 224, 2005.<br />

165. Hood, J. D. et al., Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science, 296,<br />

2404, 2002.<br />

166. Leng, Q. et al., Highly branched HK peptides are effective carriers of siRNA, J. Gene Med., 7, 977,<br />

2005.<br />

167. Pack, D. W. et al., Design and development of polymers for gene delivery, Nat. Rev. Drug Discov.,4,<br />

581, 2005.<br />

168. Kunath, K. et al., Integrin targeting using RGD-PEI conjugates for in vitro gene transfer, J. Gene<br />

Med., 5, 588, 2003.<br />

169. Erbacher, P., Remy, J. S., and Behr, J. P., Gene transfer with synthetic virus-like particles via the<br />

integrin-mediated endocytosis pathway, Gene Ther., 6, 138, 1999.<br />

170. Schiffelers, R. M. et al., Cancer siRNA therapy <strong>by</strong> tumor selective delivery with ligand-targeted<br />

sterically stabilized nanoparticles, Nucl. Acids Res., 32, e149, 2004.<br />

171. Suh, W. et al., An angiogenic, endothelial-cell-targeted polymeric gene carrier, Mol. Ther., 6, 664,<br />

2002.<br />

172. Woodle, M. C. et al., Sterically stabilized polyplex: Ligand-mediated activity, J. Control. Release,<br />

74, 309, 2001.<br />

173. Hosseinkhani, H. and Tabata, Y., PEGylation enhances tumor targeting of plasmid DNA <strong>by</strong> an<br />

artificial cationized protein with repeated RGD sequences, Pronectin, J. Control. Release, 97,<br />

157, 2004.<br />

174. Flynn, A. A. et al., A model-based approach for the optimization of radioimmunotherapy through<br />

antibody design and radionuclide selection, Cancer, 94, 1249, 2002.<br />

175. Liu, S. et al., (90)Y and (177)Lu labeling of a DOTA-conjugated vitronectin receptor antagonist<br />

useful for tumor therapy, Bioconjug. Chem., 12, 559, 2001.<br />

176. Garrison, W. M., Reaction mechanisms in radiolysis of peptides, polypeptides and proteins, Chem.<br />

Rev., 87, 381, 1987.<br />

177. Liu, S. and Edwards, D. S., Stabilization of (90)y-labeled DOTA-biomolecule conjugates using<br />

gentisic acid and ascorbic acid, Bioconjug. Chem., 12, 554, 2001.<br />

178. Mitra, A. et al., Targeting tumor angiogenic vasculature using polymer-RGD conjugates, J. Control.<br />

Release, 102, 191, 2005.<br />

179. Koivunen, E., Wang, B., and Ruoslahti, E., Phage libraries displaying cyclic peptides with different<br />

ring sizes: Ligand specificities of the RGD-directed integrins, Biotechnology (NY), 13, 265, 1995.<br />

180. Assa-Munt, N. et al., Solution structures and integrin binding activities of an RGD peptide with two<br />

isomers, Biochemistry, 40, 2373, 2001.<br />

181. Mitra, A., et al., Polymeric conjugates of mono- and bi-cyclic alphaVbeta3 binding peptides for<br />

tumor targeting, J. Control. Release, 114, 175, <strong>2006</strong>.<br />

182. Kok, R. J. et al., Preparation and functional evaluation of RGD-modified proteins as alpha(v)beta(3)<br />

integrin directed therapeutics, Bioconjug. Chem., 13, 128, 2002.<br />

183. Schraa, A. J. et al., Targeting of RGD-modified proteins to tumor vasculature: A pharmacokinetic<br />

and cellular distribution study, Int. J. Cancer, 102, 469, 2002.<br />

184. Xiong, X. B. et al., Intracellular delivery of doxorubicin with RGD-modified sterically stabilized<br />

liposomes for an improved antitumor efficacy: In vitro and in vivo, J. Pharm. Sci., 94, 1782, 2005.<br />

185. Dubey, P. K. et al., Liposomes modified with cyclic RGD peptide for tumor targeting, J. Drug<br />

Target, 12, 257, 2004.<br />

186. Bib<strong>by</strong>, D. C. et al., Pharmacokinetics and biodistribution of RGD-targeted doxorubicin-loaded<br />

nanoparticles in tumor-bearing mice, Int. J. Pharm., 293, 281, 2005.<br />

187. Seymour, L. W. et al., Effect of molecular weight (Mw) of N-(2-hydroxypropyl)methacrylamide<br />

copolymers on body distribution and rate of excretion after subcutaneous, intraperitoneal, and intravenous<br />

administration to rats, J. Biomed. Mater. Res., 21, 1341, 1987.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


184<br />

Nanotechnology for Cancer Therapy<br />

188. Seymour, L. W. et al., Influence of molecular weight on passive tumour accumulation of a soluble<br />

macromolecular drug carrier, Eur. J. Cancer, 31A, 766, 1995.<br />

189. Kissel, M. et al., Synthetic macromolecular drug carriers: Biodistribution of poly[(N-2-hydroxypropyl)methacrylamide]<br />

copolymers and their accumulation in solid rat tumors, PDA J. Pharm. Sci.<br />

Technol., 55, 191, 2001.<br />

190. Lammers, T. et al., Effect of physicochemical modification on the biodistribution and tumor accumulation<br />

of HPMA copolymers, J. Control. Release, 110, 103, 2005.<br />

191. Mitra, A. et al., Technetium-99m-Labeled N-(2-hydroxypropyl) methacrylamide copolymers:<br />

Synthesis, characterization, and in vivo biodistribution, Pharm. Res., 21, 1153, 2004.<br />

192. Ulbrich, K. et al., HPMA copolymers with pH-controlled release of doxorubicin: In vitro cytotoxicity<br />

and in vivo antitumor activity, J. Control. Release, 87, 33, 2003.<br />

193. Tucker, G. C., Alpha v integrin inhibitors and cancer therapy, Curr. Opin. Investig. Drugs, 4, 722,<br />

2003.<br />

194. Porkka, K. et al., A fragment of the HMGN2 protein homes to the nuclei of tumor cells and tumor<br />

endothelial cells in vivo, Proc. Natl. Acad. Sci. U.S.A., 99, 7444, 2002.<br />

195. Christian, S. et al., Nucleolin expressed at the cell surface is a marker of endothelial cells in<br />

angiogenic blood vessels, J. Cell Biol., 163, 871, 2003.<br />

196. Hoffman, J. A. et al., Progressive vascular changes in a transgenic mouse model of squamous cell<br />

carcinoma, Cancer Cell, 4, 383, 2003.<br />

197. Arap, W. et al., Targeting the prostate for destruction through a vascular address, Proc. Natl. Acad.<br />

Sci. U.S.A., 99, 1527, 2002.<br />

198. Joyce, J. A. et al., Stage-specific vascular markers revealed <strong>by</strong> phage display in a mouse model of<br />

pancreatic islet tumorigenesis, Cancer Cell, 4, 393, 2003.<br />

199. Ogawa, M. et al., Direct electrophilic radiofluorination of a cyclic RGD peptide for in vivo alpha(v)beta3<br />

integrin related tumor imaging, Nucl. Med. Biol., 30, 1, 2003.<br />

200. Su, Z. F. et al., In vitro and in vivo evaluation of a Technetium-99m-labeled cyclic RGD peptide as a<br />

specific marker of alpha(V)beta(3) integrin for tumor imaging, Bioconjug. Chem., 13, 561, 2002.<br />

201. Chen, X. et al., Integrin alpha v beta 3-targeted imaging of lung cancer, Neoplasia, 7, 271, 2005.<br />

202. Chen, X. et al., Synthesis and biological evaluation of dimeric RGD peptide–paclitaxel conjugate as<br />

a model for integrin-targeted drug delivery, J. Med. Chem., 48, 1098, 2005.<br />

203. Xiong, X. B. et al., Enhanced intracellular uptake of sterically stabilized liposomal Doxorubicin<br />

in vitro resulting in improved antitumor activity in vivo, Pharm. Res., 22, 933, 2005.<br />

204. Thompson, B. et al., Neutral postgrafted colloidal particles for gene delivery, Bioconjug. Chem., 16,<br />

608, 2005.<br />

205. Janssen, M. et al., Improved tumor targeting of radiolabeled RGD peptides using rapid dose fractionation,<br />

Cancer Biother. Radiopharm., 19, 399, 2004.<br />

206. Bernard, B. et al., Radiolabeled RGD-DTPA-Tyr3-octreotate for receptor-targeted radionuclide<br />

therapy, Cancer Biother. Radiopharm., 19, 173, 2004.<br />

207. Capello, A. et al., Increased cell death after therapy with an Arg-Gly-Asp-linked somatostatin analog,<br />

J. Nucl. Med., 45, 1716, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


10<br />

CONTENTS<br />

Poly(L-Glutamic Acid):<br />

Efficient Carrier of Cancer<br />

Therapeutics and Diagnostics<br />

Guodong Zhang, Edward F. Jackson, Sidney Wallace, and<br />

Chun Li<br />

10.1 Introduction ....................................................................................................................... 185<br />

10.2 Synthesis and Properties of PG......................................................................................... 186<br />

10.2.1 Chemistry ............................................................................................................ 186<br />

10.2.2 Biodegradation and Biodistribution .................................................................... 187<br />

10.3 PG-Anti-Cancer Drug Conjugates .................................................................................... 187<br />

10.3.1 Anthracyclines ..................................................................................................... 187<br />

10.3.2 Antimetabolites ................................................................................................... 188<br />

10.3.3 DNA-Binding Drugs ........................................................................................... 188<br />

10.4 PG–TXL and PG–CPT: From the Laboratory to the Clinic ............................................ 189<br />

10.4.1 PG–TXL .............................................................................................................. 189<br />

10.4.2 PG–CPT............................................................................................................... 192<br />

10.5 Combination of PG–TXL with Radiotherapy................................................................... 193<br />

10.6 PG as a Carrier of Diagnostic Agents .............................................................................. 193<br />

10.6.1 Magnetic Resonance Imaging ............................................................................. 193<br />

10.6.2 Near-Infrared Fluorescence Optical Imaging ..................................................... 195<br />

10.7 Conclusions ....................................................................................................................... 195<br />

Acknowledgments ........................................................................................................................ 196<br />

References .................................................................................................................................... 196<br />

10.1 INTRODUCTION<br />

Most anti-cancer chemotherapeutic drugs used clinically are limited <strong>by</strong> a relatively low therapeutic<br />

index, owing to toxic side effects. 1–3 Over the past several decades, two strategies of improving the<br />

therapeutic efficacy of anti-cancer agents have emerged. The first approach is the design and development<br />

of agents that can modulate the molecular processes and pathways specifically associated with<br />

tumor progression. The success of this approach is shown <strong>by</strong> the successful introduction of a new<br />

breed of molecularly targeted anti-cancer agents such as imatinib mesylate (Gleevec), gefitinib<br />

(Iressa), trastuzumab (Herceptin), and cetuximab (C225, Erbitux). Alternatively, existing<br />

anti-cancer agents can be made more effective <strong>by</strong> using delivery systems that bring more drug<br />

molecules to the tumor site when compared with conventional formulation while reducing exposure<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

185


186<br />

of normal tissues to the drug. In this context, there have been important milestones with the use of<br />

polymer–drug conjugates in their own right. In particular, poly(L-glutamic acid) (PG)–paclitaxel<br />

(TXL) (CT2103, XYOTAX) has advanced to phase III clinical trials, and PG–camptothecin (CPT)<br />

(CT2106) has been tested in phase II clinical trials. The chemistry and applications of PG and its<br />

conjugates with various chemotherapeutic agents were previously reviewed. 1 In this chapter, the<br />

applications of PG in the delivery of both chemotherapeutic and diagnostic agents will be updated.<br />

The results of clinical studies of XYOTAX and CT2106 will also be summarized. Previous chapters<br />

in this book discuss, in depth, the rationale for the use of polymeric conjugates as cancer therapeutics.<br />

10.2 SYNTHESIS AND PROPERTIES OF PG<br />

PG anti-cancer drug conjugates consist of three parts: the PG backbone, the anti-cancer drug, and<br />

spacers that link the drug molecules with the PG backbone. Both the properties of PG (i.e., the<br />

molecular weight and its distribution) and the selection of spacers can directly affect the pharmacokinetics<br />

at both the whole-body and cellular level. For this reason, the synthetic chemistry of the<br />

preparation of PG will first be discussed. A summary of the physicochemical properties of PG<br />

conjugates will follow.<br />

10.2.1 CHEMISTRY<br />

Nanotechnology for Cancer Therapy<br />

PG is usually prepared <strong>by</strong> removing the benzyl protecting group of poly(g-benzyl-L-glutamate) that<br />

is attained <strong>by</strong> polymerization of g-benzyl-L-glutamate N-carboxyanhydride (NCA) monomer. 4<br />

Direct phosgenation of g-benzyl-L-glutamate produces the corresponding NCA monomer. 5,6 In<br />

this reaction, g-benzyl-L-glutamate is suspended in a dry inert solvent such as ethyl acetate,<br />

dioxane, or tetrahydrofuran, and it is allowed to heterogeneously react with the cyclizing reagent<br />

that is normally triphosgene. Researchers have used both the protic and aprotic initiators in the<br />

polymerization of g-benzyl-L-glutamate NCA monomer. 7 Protic initiators such as primary amines<br />

are acylated <strong>by</strong> attack on the 5-position of the NCA, whereas aprotic initiators such as tertiary<br />

amines and alkoxides act as general bases. The ring-opening polymerization of NCA initiated <strong>by</strong><br />

amines usually produces polymers with relatively broad molecular weight distributions. To circumvent<br />

this problem, Deming 8 developed the zero-valence nickel catalyst bipyNi(COD) (bipy:<br />

2,2 0 -bipyridyl; COD: 1,5-cyclooctadiene) to initiate polymerization of NCAs. The active sites<br />

derived with this initiator are less accessible to side reactions when compared with that derived<br />

with amine initiators because of steric and electronic effects, resulting in polymers with a narrow<br />

molecular weight distribution (Mw/Mn, 1.05–1.15).<br />

Impurities in NCAs can have a detrimental effect on the reproducibility of polyamino acid<br />

synthesis. NCAs are usually subjected to repeat recrystallization to eliminate trace amounts of<br />

impurities. A recent report described an efficient method of removal of hydrogen chloride and the<br />

hydrochloride salt of unreacted starting amino acids. 9 This method consists of extraction of NCA<br />

solution in ethyl acetate with water and an aqueous alkali solution at 08C. Using highly purified<br />

monomer NCAs, Aliferis et al. 10 were able to achieve living polymerization initiated with primary<br />

amine initiators. This technique produced poly(g-benzyl-L-glutamate) with an average molecular<br />

weight as high as 1.66!10 5 Da and with a relatively narrow molecular weight distribution (Mw/<br />

Mn, 1.40).<br />

The benzyl protecting group in poly(g-benzyl-L-glutamate) is removed <strong>by</strong> treatment with<br />

hydrogen bromide. Researchers have also used alkaline hydrolysis of poly(L-methyl glutamate)<br />

to prepare PG; however, racemization occurred during this process. 11 To avoid using harsh conditions<br />

for the removal of protecting groups, investigators have prepared glutamic acid NCAs<br />

having ester protecting groups that are labile under mild acidic conditions. For example, PG can<br />

be readily obtained from poly(g-piperonyl-L-glutamate) <strong>by</strong> treating the polymer with trifluoroacetic<br />

acid. 12<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 187<br />

PG-based block copolymers can be prepared using macroinitiators or polymer coupling reactions.<br />

The block copolymer PG–vinylsulfone–polyethylene glycol (PEG) was prepared <strong>by</strong> directly reacting a<br />

heterofunctional PEG, vinylsulfone–PEG–N-hydroxysuccinimide (NHS), with an amino group at one<br />

end of the PG chain. 13 Also, Nishiyama et al. 14 used the macroinitiator PEG–NH 2 to initiate polymerization<br />

of g-benzyl-L-glutamate NCA to produce a PEG–PG block copolymer.<br />

Another polymer derived from poly(g-benzyl-L-glutamate) is poly(hydroxyethylglutamate)<br />

(PHEG) that is also a suitable candidate drug carrier. 15 PHEG is a water-soluble, neutral<br />

polymer that can be readily prepared from poly(g-benzyl glutamate) via aminolysis with<br />

2-aminoethanol using 2-hydroxypyridine as a catalyst. 16<br />

The linkage of polymer–drug conjugates is important to the in vivo release of a drug that should<br />

be stable during circulation but should also allow drug release at an appropriate rate at the tumor<br />

tissue. 17 Generally, two kinds of linkage are involved in anti-cancer drug delivery <strong>by</strong> using<br />

polymer–drug conjugates: enzymatic hydrolysis and nonenzymatic hydrolyzable linkage. Many<br />

proteolytic enzymes such as cysteine proteinase cathepsins are expressed on the surface of metastatic<br />

tumor cells. 18,19 The linker chemistry can be designed in such a way that the active agents<br />

would be released from the polymer–drug conjugates only upon exposure to the proteinases in the<br />

tumors. Alternatively, hydrolytically labile, pH-sensitive linkages (i.e., hydrozone, ester, acetals,<br />

etc.) have exhibited their utility for tumorotropic or lysosomotropic delivery because the microenvironments<br />

of solid tumors are known to be acidic. 20,21 PG–TXL conjugate releases TXL through<br />

both mechanisms: backbone degradation mediated through proteolytic enzymes and side chain<br />

hydrolysis of the ester bond formed between glutamic acid and TXL.<br />

10.2.2 BIODEGRADATION AND BIODISTRIBUTION<br />

The enzymatic degradability of PG–drug conjugates is influenced <strong>by</strong> their structure, composition,<br />

and charge as well as the physicochemical properties of the drug attached to PG. 22–25 PG is more<br />

susceptible to lysosomal degradation than are poly(aspartic acid) and poly(D-glutamic acid). 26<br />

Cysteine proteases, particularly cathepsin B, play key roles in the lysosomal degradation of<br />

PG. 22,27 Although researchers have identified oligomeric glutamic acids as the primary degradation<br />

products of PG, 28 more recent results demonstrated that monomeric L-glutamic acid is produced in<br />

the lysosomal degradation of PG. 29 Degradation of the PG backbone may not necessarily lead to the<br />

release of free drug in every case.<br />

The biodistribution of PG and its drug conjugates depends on the molecular weight of PG.<br />

Polymers with a molecular weight lower than the renal clearance threshold are rapidly removed<br />

from the blood circulation <strong>by</strong> glomerular filtration. McCormick-Thomson and Duncan 27 found that<br />

the biodistribution of copolymers of glutamic acid and other hydrophobic amino acids was markedly<br />

influenced <strong>by</strong> the composition of the copolymers. These results suggest that the biodistribution<br />

of PG–drug conjugates is a function of the drugs used, the degree of modification, and the molecular<br />

weight of PG.<br />

10.3 PG-ANTI-CANCER DRUG CONJUGATES<br />

Since doxorubicin (Adriamycin [ADR]) was conjugated with PG in the 1980s, 30 investigators have<br />

studied various anti-cancer drugs such as anthracyclines, antimetabolites, DNA-binding drugs,<br />

TXL, and CPT. In particular, researchers have evaluated PG–TXL in phase III clinical trials and<br />

PG–CPTin phase II clinical trials.<br />

10.3.1 ANTHRACYCLINES<br />

The anthracycline group has been one of the most extensively studied groups of drugs used in<br />

polymer–drug conjugate delivery systems, especially in the 1980s. Based on the assumption that<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


188<br />

greater selectivity toward tumors over normal tissues can be achieved if PG–ADR conjugates only<br />

degrade and release ADR after being endocytosed <strong>by</strong> tumor cells, Van Heeswijk and colleagues 30<br />

investigated three different PG conjugates containing ADR where the amine group in ADR was<br />

bound to the side chain carboxyl groups of high-molecular-weight PG either directly (with an amide<br />

bond) or indirectly through GlyGly and GlyGlyGlyLeu spacers, respectively. They investigated the<br />

degradability of the conjugates mediated <strong>by</strong> lysosomal enzymes and the subsequent release of ADR<br />

or ADR-peptide products with low molecular weights using reverse-phase high-performance liquid<br />

chromatography. The total amount of ADR released after 77 h of incubation was 3.6% for PG-Gly-<br />

GlyGlyLeu-ADR; 1.0% for PG-GlyGly-ADR; and 0.5% for PG–ADR, suggesting the importance<br />

of introducing an enzymatically degradable spacer between PG and the drug. In vitro, these<br />

conjugates exhibited reduced cytotoxicity against L1210 leukemia cells when compared with<br />

free ADR. In vivo, animals treated with the polymer conjugate containing the tetrapeptide<br />

spacer showed a similar mean survival duration as compared to those treated with free ADR,<br />

whereas the conjugate without a peptide spacer was completely inactive. 28,31<br />

Some have used hydrolytically labile ester bonds 32 and hydrozone bonds 33 to couple doxorubicin<br />

or daunorubicin (Dau) with PG. In these studies, the ester bond was formed <strong>by</strong> a reaction of<br />

14-bromo-daunorubicin with the carboxylic group of PG via a nucleophilic substitution reaction in<br />

an alkaline aqueous medium. With intravenous administration of these conjugates into mice<br />

bearing MS-2 sarcoma or Gross’ leukemia, these investigators found that the drug conjugates’<br />

potency and efficacy were correlated with the molecular weight of the carrier. For example,<br />

anti-tumor activity improved when the molecular weight increased from 14,000 to 60,000 Da at<br />

an equivalent doxorubicin dose of 30 mg/kg. These results can be attributed to the increased<br />

circulation times of conjugates with high molecular weights. Condensing the methylketone in<br />

Dau with hydrazide-derived PG yielded PG-hydrazone–Dau conjugates. The acid-sensitive conjugates<br />

were less cytotoxic than free Dau was in vitro against mouse lymphoma cells. However, these<br />

conjugates showed significant anti-tumor activity when intravenously injected into mice bearing<br />

intraperitoneally inoculated Yac lymphoma. 33<br />

10.3.2 ANTIMETABOLITES<br />

To enhance the efficacy of 1-b-D-arabinofuranosylcytosine (ara-C) in simple dosage schedules,<br />

researchers synthesized two ara-C conjugates with PG via an amide bond: one where ara-C is<br />

directly coupled with the N-4 of ara-C to the carboxyl groups of PG and one where ara-C is linked<br />

with PG via the aminoalkylphosphoryl side chain introduced at the C-5 0 of ara-C. 34 Both conjugates<br />

exhibited markedly decreased cytotoxicity in L1210 murine leukemia cells when compared with<br />

free ara-C. However, the anti-tumor activity of both conjugates was greater than or equal to that of<br />

free ara-C in mice bearing L1210 tumors after a single intraperitoneal injection of the conjugate.<br />

The authors suggested that both slow cleavage of free ara-C from the conjugates and protection of<br />

ara-C from deactivation <strong>by</strong> cytidine deaminase contributed to the enhanced anti-tumor activity of<br />

PG–ara-C conjugates in vivo.<br />

10.3.3 DNA-BINDING DRUGS<br />

Nanotechnology for Cancer Therapy<br />

Investigators have conjugated several DNA-binding drugs, including cyclophosphamide, 35<br />

L-phenylalanine mustard (melphalan), 36 mitomycin C (MMC), 37–39 and cis-dichlorodiammineplatinum<br />

(II) (cisplatin [CDDP]) 40,41 with PG. Moromoto et al. 36 prepared PG–melphalan <strong>by</strong> coupling<br />

the amine group of melphalan with the side-chain carboxyl group of PG in the presence of a watersoluble<br />

carbodiimide. 36 They investigated the anti-tumor activity of this conjugate in rats bearing<br />

sarcoma induced <strong>by</strong> subcutaneous inoculation of sarcoma cells. The authors used PG– 3 H-phenylalanine<br />

as a model compound to demonstrate that free drug could be released from the conjugate<br />

and that the conjugate had a tendency to be absorbed through lymphatic routes when compared with<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 189<br />

free 3 H-phenylalanine after subcutaneous administration. 36 Roos et al. 37 conjugated MMC with PG<br />

<strong>by</strong> using the aziridine amine of MMC. 37,38 They found that the release rate, in vitro cytotoxicity,<br />

and in vivo anti-tumor activity of PG–MMC conjugates were influenced <strong>by</strong> the extent of MMC<br />

substitution. Furthermore, Seymour et al. 42 conjugated MMC with PHEG using an oligopeptide<br />

spacer and investigated the effect of this oligopeptide structure on the rate of drug release to<br />

optimize the MMC–oligopeptide–PHEG conjugates. The rate of MMC release from the conjugates<br />

was affected <strong>by</strong> the structure of the oligopeptide spacer; spacers bearing a terminal Gly had the<br />

fastest release of MMC when compared with other peptide spacers. Another study showed a<br />

correlation between the in vitro cytotoxicity of PG–MMC conjugates against B16F10 melanoma<br />

and C26 colorectal carcinoma cells and their hydrolytic stability. 38 Significant in vivo activity of<br />

the conjugates in mice bearing P338 leukemia or C26 colorectal carcinoma seems to result from<br />

both the stability of the conjugates in the blood and the rapid drug release, owing to combined<br />

chemical and enzymatic hydrolysis.<br />

The clinical use of CDDP is limited <strong>by</strong> significantly toxic side effects such as acute nephrotoxicity<br />

and chronic neurotoxicity and a low therapeutic index. To reduce its toxicity and to<br />

maintain prolonged drug activity, investigators complexed CDDP with PG. 40,41 They found that<br />

PG–CDDP complex (60 mol CDDP/mol PG; molecular weight, 40,000 Da) was more thermodynamically<br />

stable and has reduced systemic toxicity when compared with CDDP. PG–CDDP<br />

was effective in suppressing the growth of OVCAR-3 human ovarian carcinoma in athymic<br />

mice and showed a broader therapeutic dose range (80% survival at 3–12 mg/kg) when compared<br />

with the narrow, inconsistent effective dose range of free CDDP (80% survival at 1.0–2.5 mg/kg),<br />

suggesting that the therapeutic index of CDDP can be improved <strong>by</strong> complexing it with PG.<br />

10.4 PG–TXL AND PG–CPT: FROM THE LABORATORY TO THE CLINIC<br />

10.4.1 PG–TXL<br />

Taxanes are a class of the widely used and clinically active cytotoxic agents that exert their action<br />

<strong>by</strong> promoting tubulin polymerization and microtubule assembly. 43 Because of its poor aqueous<br />

solubility, use of TXL requires Cremophor and ethanol mixture as a vehicle, and it must be infused<br />

over 3–24 h. The conjugate PG–TXL where TXL is attached to PG via ester bonds has demonstrated<br />

significantly enhanced anti-tumor efficacy and improved safety when compared with TXL in<br />

preclinical studies. 44<br />

The release of TXL from PG–TXL and the metabolism of PG–TXL in both in vitro and in vivo<br />

models have been investigated. 45 These studies showed that when PG–TXL was incubated in<br />

buffered saline or plasma (mouse or human) for 24 h at 378C, less than 14% of the bound TXL<br />

was released, suggesting that PG–TXL is relatively resistant to plasma esterase.<br />

Tumor uptake of TXL and PG–TXL was compared using [ 3 H]TXL and PG–[ 3 H]TXL. 46 When<br />

free [ 3 H]TXL was incubated with MDA-MB453 cells, [ 3 H]TXL was taken up rapidly <strong>by</strong> the cells<br />

and was followed <strong>by</strong> a rapid drug efflux process. In contrast, when the tumor cells were incubated<br />

with PG–[ 3 H]TXL, a slower uptake and more persistent retention of radioactivity was observed.<br />

These data suggest that a fraction of PG–[ 3 H]TXL was taken up <strong>by</strong> tumor cells, possibly through<br />

pinocytosis, and that [ 3 H]TXL was more readily pumped out of the cells than were the more<br />

hydrophilic PG–[ 3 H]TXL and its degradation products. A recent report confirmed monoglutamyl-2<br />

0 -TXL and diglutamyl-2 0 -TXL as the major intracellular metabolites of PG–TXL. 47<br />

Hydrolysis of these metabolites led to release of free TXL. Specific enzyme inhibitors such as<br />

CA-074 methyl ester, a cell-permeable irreversible inhibitor of cathepsin B, and EST, a cellpermeable<br />

irreversible inhibitor of cysteine protease, decreased the formation of monoglutamate<br />

TXL and free TXL in a tumor cell line that had been incubated with PG–TXL. All of these findings<br />

are consistent with the results of proteolysis of the PG backbone through the action of cellular<br />

dipeptidases. Another experiment showed that the metabolism of PG–TXL in non tumor-bearing<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


190<br />

Nanotechnology for Cancer Therapy<br />

cathepsin B homozygous knockout mice is reduced, but not eliminated; this seems to confirm that<br />

in addition to cathepsin B, other cysteine proteases on the surface of tumor cells play important<br />

roles in the release of TXL from PG–TXL. 47<br />

PG–TXL’s anti-tumor activity was first assessed in a variety of syngeneic and xenogeneic<br />

tumor models. For example, the maximum tolerated dose of PG–TXL after a single intravenous<br />

injection in rats and mice was 60 and 160 mg/kg, respectively. 44 In comparison, the maximum<br />

tolerated dose of TXL in rats and mice was 20 and 60 mg/kg, respectively. Therefore, PG–TXL’s<br />

use represented a twofold and threefold improvement in toxicity in rats and mice, respectively. To<br />

determine if PG–TXL has a broad spectrum of anti-tumor activity, its therapeutic activity was<br />

evaluated against four syngeneic murine tumor cell lines (MCa-4 breast carcinoma, MCa-35 breast<br />

carcinoma, HCa-1 hepatocarcinoma, and FSa-II sarcoma) intramuscularly inoculated into C3Hf/<br />

Kam mice. 48 The anti-tumor and antimetastatic activities of PG–TXL were investigated using<br />

intraperitoneal injection of the SKOV3ip1 human ovarian tumor cell line in nude mice 49 and<br />

human MDA-MB-435-Lung2 breast tumors grown in the mammary fat pads of nude mice. 48<br />

Treatment with PG–TXL exhibited significantly better anti-tumor activity than did treatment<br />

with TXL alone in all of the tumor models.<br />

The pharmacokinetic profile and tissue distribution of PG–TXL were examined to verify the<br />

enhanced permeability and retention effect of macromolecules. 45 It was found that PG–[ 3 H]TXL<br />

has a much longer half-life in plasma (317 min) than [ 3 H]TXL does (29 min). Consequently, the<br />

area under the tissue-concentration-time curve (AUC) in tumors was five times greater when mice<br />

were injected with PG–TXL than when they were injected with TXL. Therefore, enhanced tumor<br />

uptake and sustained release of TXL from PG–TXL in tumor tissue seem to be among one of the<br />

major factors contributing to PG–TXL’s markedly improved anti-tumor activity in vivo.<br />

In the XYOTAX formulation used in clinical studies, the median molecular weight of PG–TXL<br />

is 48,000 Da, and the content of TXL is about 37% <strong>by</strong> weight, equivalent to approximately one<br />

TXL molecule for every 11 glutamic acid units in each PG polymer chain. 47 This formulation<br />

eliminated the use of Cremophor and alcohol, and it allows infusion of TXL over 30 min. In initial<br />

clinical trials in the United Kingdom, the investigators gave PG–TXL (CT-2103) to cancer patients<br />

in a 30-min infusion every three weeks at doses ranging from 30 to 720 mg/m 2 . 50 They detected<br />

CT-2103 in all the patients’ plasma and observed a long plasma half-life of up to 185 h. Importantly,<br />

the peak plasma concentration of released free TXL was less than 0.1 mM 24hafter<br />

administration of CT-2103 at doses up to 480 mg/m 2 (176 mg/m 2 TXL equivalent). As shown in<br />

Figure 10.1, the plasma concentrations of PG–TXL biphasically declined. The distribution phase<br />

was prolonged, and the apparent monoexponential terminal phase associated with drug elimination<br />

appeared approximately 48 h after administration of PG–TXL. The plasma concentration of<br />

PG–TXL in the terminal phase declined slowly, and the drug could be detected in plasma three<br />

weeks after administration at a dose of 200 mg/m 2 . In comparison, the plasma concentration of free<br />

TXL paralleled with the PG–TXL concentration. Importantly, this study found that the AUC of<br />

free TXL was about 1–2% of the AUC of PG–TXL that supported the in vivo stability of PG–TXL<br />

in the plasma and the slow, prolonged release of the active moiety. 51 The steady-state volume of<br />

distribution ranged from 1.3 to 5.9 l/m 2 and was low that suggested that the distribution of PG–TXL<br />

was restricted mainly to the plasma and other extracellular body fluids. 47<br />

More than 400 patients have received PG–TXL in phase I and II trials. 52,53 Drug-related events<br />

that were reported in 10–20% of the patients included thrombocytopenia, diarrhea, leukopenia,<br />

myalgia, arthralgia, and anemia. Compared with conventional TXL-based treatment, PG–TXL<br />

showed three safety-related advantages. First, alopecia was rare, and complete hair loss was not<br />

observed. Second, nausea and vomiting were uncommon. Third, hypersensitivity reactions were<br />

rarely observed, and those that did occur were usually mild to moderate; therefore, routinely used<br />

prophylactic premedications were not required. The incidence of significant hypersensitivity<br />

reactions was less than 1% with no grade 4 hypersensitivity reactions.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 191<br />

Plasma concentration (µg/ml)<br />

10 6<br />

10 5<br />

10 4<br />

10 3<br />

10 2<br />

10<br />

1<br />

(n = 15)<br />

(n = 15)<br />

(n = 15)<br />

(n = 15)<br />

(n = 13)<br />

Free paclitaxel<br />

0 5 10<br />

Average dose 496 mg<br />

Conjugates taxanes<br />

(n = 14)<br />

(n = 12)<br />

Phase II trials CT-2103, including the multicenter open-label studies (CTI-1071 and CTI-<br />

1069), have been completed. 47,53 The former study aimed to determine the rate of response and<br />

time to disease progression in a heterogeneous population of patients with advanced epithelial<br />

ovarian cancer who received CT-2103. 47,53 Ninety-nine patients registered in this trial received<br />

PG–TXL at a dose of 175 mg/m 2 every three weeks. The toxic effects were mild in this heavily<br />

pretreated population and consisted of the following: grade 3 neuropathy (nZ15), grade 3 neutropenia<br />

(nZ10), and grade 4 neutropenia (nZ4). Of 18 patients who received one or two prior<br />

chemotherapy regimens and had CDDP-sensitive disease, five (28%) had a response, and six (33%)<br />

had stable disease (SD). The cancer was difficult to treat in 21 patients with platinum-resistant<br />

disease who underwent pretreatment, yet the investigators observed responses in two patients<br />

(10%) and SD in four patients (19%). In CTI-1069, 47 the researchers aimed to evaluate the efficacy<br />

and tolerability of PG–TXL in patients with nonsmall cell lung cancer who were 70 years of age or<br />

older and had an Eastern Cooperative Oncology <strong>Group</strong> (ECOG) performance status of 2 (PS2).<br />

Thirty patients were registered in this trial, and they received PG–TXL at a dose of 175 mg/m 2<br />

(nZ28) or 235 mg/m 2 (nZ2) every three weeks. Two patients had a partial response, whereas 16<br />

patients had SD lasting at least 10 weeks. Among the 28 patients who received PG–TXL at 175 mg/<br />

m 2 , the median survival duration was 8.1 months in those with an ECOG performance status of 0<br />

(PS0) or 1 (PS1) and 5.4 months in PS2 patients. Both of the patients who received PG–TXL at<br />

235 mg/m 2 died within 30 days after treatment; one died of neutropenia, whereas the other died of<br />

septic shock and renal failure.<br />

More than 700 patients with lung cancer have participated in phase III trials<br />

PG–TXL(XYOTAX) that include two phase III trials of XYOTAX as first-line treatment in PS2<br />

patients (STELLAR 3 and STELLAR 4) and one phase III trial of XYOTAX as second-line<br />

treatment in PS0, PS1, and PS2 patients (STELLAR 2). 47 The total number of patients in the<br />

STELLAR 2, STELLAR 3, and STELLAR 4 trials was 850, 400, and 477, respectively. Cell<br />

Therapeutics Inc. (CTI) announced the results of these trials in March and May 2005. Although<br />

none of the three trials met their primary end points, they did demonstrate similar efficacy, reduced<br />

Day<br />

15<br />

(n = 8)<br />

20 25<br />

FIGURE 10.1 Plasma pharmacokinetics of PG–TXL and free TXL in cancer patients. The data are from four<br />

phase I dose-escalation studies using 1-, 2-, and 3-week schedules. In all of the studies, PG–TXL was<br />

administrated in a short intravenous infusion. (From Singer, J. W., Shaffer, S., and Baker, B., Anti-cancer<br />

Drugs, 16, 243–254, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


192<br />

side effects, and more convenient administration of XYOTAX when compared with the control<br />

drugs (TXL, docetaxel, gemcitabine, and vinorelbine).<br />

In addition, clinical data from a pooled analysis of CTI’s STELLAR 3 and 4 trials showed that<br />

in the 198 women treated on those trials, superior survival was observed in those who received<br />

XYOTAX (pZ0.03). The most notable impact was among women less than 55 years old and<br />

presumably pre-menopausal who were treated with XYOTAX compared to standard chemotherapy<br />

(median survival 10.0 vs. 5.3 months, hazard ratioZ0.51, log rank pZ0.038) while a survival<br />

trend (pZ0.134) was observed in women 55 years of age and older (post-menopausal) (http://<br />

www.ctiseattle.com). The favorable anti-tumor activity among women less than 55 years of age<br />

stimulated the initiation of the PIONEER 1 clinical trial where about 600 PS2 chemotherapy-naïve<br />

women with advanced stage NSCLC will be recruited. Each study arm of approximately 300<br />

patients will be randomized to receive either XYOTAX at 175 mg/m 2 or TXL at 175 mg/m 2<br />

once every three weeks. The primary endpoint is superior overall survival with several secondary<br />

endpoints including disease control, response rate in patients with measurable disease, time to<br />

disease progression, and disease-related symptoms.<br />

CTI also reported the preliminary results of a phase II study of XYOTAX in combination with<br />

carboplatin for first-line induction and single-agent maintenance therapy for advanced-stage III/IV<br />

ovarian cancer in May 2005. Among the 82 patients in this study, 98% had a major tumor response,<br />

85% had a complete response, and 12% had a partial response during induction of the therapy. At a<br />

dose of 175 mg/m 2 , XYOTAX and carboplatin (AUCZ6) had grade 3/4 side effects, including<br />

thrombocytopenia (55%), neuropathy (23%), febrile neutropenia (19%), nausea (15%), anemia<br />

(11%), and vomiting (7%). The investigators found no grade 4 neuropathy, and only 4% patients<br />

needed dose delay because of neutropenia. CTI and the Gynecologic Oncology <strong>Group</strong> are collaborating<br />

on phase III studies for which they expect to enroll about 1550 patients to receive<br />

treatment with carboplatin and TXL initially followed <strong>by</strong> XYOTAX for consolidation in half of<br />

the patients.<br />

10.4.2 PG–CPT<br />

Nanotechnology for Cancer Therapy<br />

The CPTs are a family of synthetic and semisynthetic analogues of 20(S)-CPT that exhibit a broad<br />

range of anti-cancer activity <strong>by</strong> inhibiting topoisomerase-1 activity. Two properties of CPT<br />

compounds limit their therapeutic efficacy in humans: instability of the lactone form because of<br />

preferential binding of the carboxylate to serum albumin and a lack of aqueous solubility. PG is an<br />

effective solubilizing carrier of CPT and serves to protect the E-ring lactone structure in CPT. The<br />

PG–CPT conjugate was prepared <strong>by</strong> directly coupling the hydroxy group at the C20(S)-position of<br />

CPT with the carboxylic acid of PG. 54 When given intravenously in four doses every four days at an<br />

equivalent CPT dose of 40 mg/kg, PG–CPT delayed the growth of established H322 human lung<br />

tumors subcutaneously grown in nude mice. In mice that received intratracheal inoculation of H322<br />

cells, the same treatment prolonged the median survival duration in mice <strong>by</strong> fourfold when<br />

compared with that in untreated control mice. These results showed that PG was as efficient<br />

carrier of CPT.<br />

Studies have systematically investigated the structural effects of the anti-tumor efficacy of CPT,<br />

including linkers between PG and CPT, the point of attachment of PG on the CPT molecule, the<br />

polymer molecular weight, and drug loading. 55–57 First, coupling through the 20(S)-hydroxy group<br />

of CPT with or without a glycine linker yielded the most active conjugates because this site is<br />

located in close proximity to the lactone ring; therefore, the E-ring lactone is better protected <strong>by</strong> the<br />

linked PG chains. Second, increasing the molecular weight of PG from 33 to 50 kDa improved the<br />

anti-tumor efficacy of PG–CPT, probably because of an increased plasma half-life and reduced<br />

renal clearance. Third, based on the CPT-equivalent dosing levels, the investigators compared<br />

various linkers and found that PG-Gly-CPT and PG–(4-O-butyryl)–CPT seemed had the highest<br />

anti-tumor activity. Moreover, the linkers also affected the maximum drug payload; for example,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 193<br />

only 15% (<strong>by</strong> weight) of CPT loading could be achieved for direct conjugation of PG–CPT <strong>by</strong> the<br />

ester linkage because of steric hindrance. On the other hand, up to 50% CPT loading could be<br />

achieved for the case with a glycine as linker. Fourth, the researchers observed improved anti-tumor<br />

efficacy of PG-Gly-CPT against HT-29 colon cancer and NCI-H460 lung carcinoma with increased<br />

CPT loading. However, increasing the loading of CPT to 47% resulted in significantly reduced<br />

solubility. Therefore, they recommended further investigation of PG-Gly-CPT with loading of CPT<br />

at 30–35%.<br />

10.5 COMBINATION OF PG–TXL WITH RADIOTHERAPY<br />

Combining chemotherapy and radiotherapy has significantly improved response and survival rates<br />

in patients with many solid tumors. Many chemotherapeutic agents can increase the radiosensitivity<br />

of tumors, potentiating the tumor response to radiation-caused damage. It was hypothesized that<br />

combining radiotherapy and chemotherapy using a polymer–drug conjugate may lead to a stronger<br />

radiosensitizing effect than using the drug alone. Irradiation can, in turn, potentiate the tumor<br />

response to polymer–drug conjugates <strong>by</strong> increasing tumor vascular permeability and the uptake<br />

of these conjugates into solid tumors. To test this hypothesis, PG–TXL was used as a model<br />

polymer–drug conjugate. 58 Administration of PG–TXL delayed the growth of OCa-1 syngeneic<br />

murine ovarian tumors in C3Hf/Kam mice. However, when PG–TXL was given in combination<br />

with tumor irradiation, significantly enhanced anti-tumor activity was observed. Using tumor<br />

growth delay as an end point, enhancement factors ranging from 1.36 to 4.4 were observed;<br />

these values depended on the doses of PG–TXL and radiation delivered. It was found that complete<br />

tumor regression occurred with the use of increased radiation doses (O10 Gy) and PG–TXL doses<br />

(O80 mg/kg equivalent TXL). 59 Similar results were observed in a mammary MCa-4 carcinoma<br />

model. 60 In contrast, it was found that combined radiotherapy and TXL treatment yielded an<br />

enhancement factor of less than 1.0 in MCa-4 tumors, indicating that conjugation of TXL with<br />

PG is necessary to improve that radiosensitization effect of TXL. When the treatment end point was<br />

tumor cure, enhancement factors as high as 8.4 and 7.2 were observed after fractionated and singledose<br />

radiotherapy, respectively. 61,62<br />

To determine if prior irradiation affects tumor uptake of PG–TXL, [ 3 H]PG–TXL was injected<br />

into mice with OCa-1 ovarian tumors 24 h after local irradiation at 15 Gy. 59 The uptake of<br />

[ 3 H]PG–TXL in irradiated tumors was 28–38% higher than that in nonirradiated tumors at different<br />

times after injection of [ 3 H]PG–TXL, suggesting that irradiation increased the accumulation of<br />

PG–TXL in the tumors. Therefore, the super-synergistic effect of combined radiotherapy and<br />

PG–TXL-based chemotherapy is partly ascribed to the enhanced permeability and retention<br />

effect of macromolecules caused <strong>by</strong> irradiation.<br />

10.6 PG AS A CARRIER OF DIAGNOSTIC AGENTS<br />

10.6.1 MAGNETIC RESONANCE IMAGING<br />

Noninvasive imaging of intravascular compartments is of critical importance in the clinic. Many<br />

diseases such as infections, ulcers, cardiovascular diseases, and solid tumors involve gross hemorrhaging,<br />

abnormal vascular growth, and/or vascular occlusion. 63,64 Magnetic resonance imaging<br />

(MRI) with blood-pool contrast agents can be used to perform minimally invasive angiography,<br />

assess angiogenesis, quantify and measure the spacing of blood vessels, and measure blood volume<br />

and flow. 63,64 MRI blood-pool contrast agents are often paramagnetic gadolinium (Gd) chelates of<br />

high-molecular-weight polymers that are largely retained within the intravascular space during<br />

MRI. 65 Recent studies showed that tumor vascular permeability measurements obtained using<br />

polymeric contrast agents enhanced dynamic MRI that correlated with tumor microvessel<br />

density counts, suggesting that MRI can be used to characterize tumor angiogenic activity. 66,67<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


194<br />

Nanotechnology for Cancer Therapy<br />

Furthermore, MRI with polymeric contrast agents may permit more accurate grading of tumor<br />

invasiveness than those with smaller molecular weight contrast agents, such as Gd–diethylenetriaminepentaacetic<br />

acid (DTPA). 68,69<br />

Investigators have synthesized various polymeric contrast agents for MRI, including human<br />

serum albumin, 70 polylysine, 71,72 dextran, 73,74 dendrimers, 75–78 polyamide, 74,79 and grafted copolymers,<br />

80 and they evaluated them as blood-pool imaging agents. Although most of these agents<br />

fulfill the criteria for long blood circulation time and high MRI relaxivity, their safety remains<br />

unestablished. The clinical applications of many current macromolecular contrast agents are<br />

limited <strong>by</strong> slow excretion from the body and the potential toxicity of free Gd 3C ions released <strong>by</strong><br />

the metabolism of the contrast agents. 71–73,81,82 For albumin-based products, the possibility of<br />

immunogenic responses makes the use of serum proteins less attractive. Dendrimer-based bloodpool<br />

agents have the advantage of extremely narrow molecular weight distributions; however, these<br />

agents are not biodegradable.<br />

Ideally, polymeric contrast agents are degraded and cleared from the body after completion of<br />

MRI. Based on its good biocompatibility and biodegradability, the metal chelator DTPA was<br />

conjugated with PG, and the physicochemical and imaging properties of the resulting polymeric<br />

contrast agents were evaluated. 83 One of these agents, PG–Bz–DTPA–Gd, was synthesized from<br />

PG and monofunctional p-aminobenzyl-DTPA (penta-tert-butyl ester). It was found that PG–Bz–<br />

DTPA–Gd was readily degraded upon exposure to an aqueous buffered solution containing cathepsin<br />

B. The T1 relaxivity of PG–Bz–DTPA–Gd at 1.5 T was four times greater than that of smallmolecular-weight<br />

Gd–DTPA. Figure 10.2 compares the parametric AUC magnetic resonance<br />

images for signal intensity integrated over 90 s and 10 min after intravenous injection of gadopentetate<br />

dimeglumine (Magnevist; Gd–DTPA) and PG–Bz–DTPA–Gd. Magnevist (743 Da) is a<br />

contrast agent that is clinically used. Whereas Magnevist rapidly diffused into the extravascular<br />

fluid space over 90 s, PG–Bz–DTPA–Gd was largely retained in the blood vessels for up to 10 min.<br />

Indeed, contrast enhancement of the vascular compartments was still visible 2 h after injection of<br />

PG–Bz–DTPA–Gd. 83<br />

In an effort to increase the rate of polymer degradation, Lu et al. 84 prepared PG–cystamine–<br />

[Gd(III)-1,4,7,10-tetraazacyclododecane-1,4,7,10-tetraacetic acid (DOTA)] (molecular weight of<br />

PG, 50,000 Da) where the metal chelator DOTA was conjugated with PG using a disulfide bond.<br />

They showed that glutathione and other endogenous sulfhydryl-containing biomolecules could<br />

exchange their free SH group with an S–S bond in PG–cystamine–[Gd(III)–DOTA], resulting in<br />

FIGURE 10.2 (See color insert following page 522.) Comparison of parametric AUC magnetic resonance<br />

images obtained at 90 s and 10 min after intravenous injection of Magnevist and PG–Bz–DTPA–Gd. Arrow:<br />

Blood vessel.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 195<br />

PG<br />

PG<br />

Cl<br />

rapid release of Gd(III)–DOTA from the polymer and subsequent clearance of Gd-containing<br />

species from the body. Use of PG–cystamine–[Gd(III)–DOTA] produced significant blood-pool<br />

contrast enhancement on MRI scans of the heart and blood vessels in nude mice bearing OVCAR-3<br />

human ovarian carcinoma xenograft when compared with use of small-molecular-weight<br />

contrast agents.<br />

10.6.2 NEAR-INFRARED FLUORESCENCE OPTICAL IMAGING<br />

Near-infrared (NIR) optical imaging has unique advantages for diagnostic imaging of solid tumors.<br />

Specifically, NIR imaging is a potentially safe, noninvasive method of detecting solid tumors as<br />

NIR light (650- to 900-nm wavelengths) can penetrate several centimeters into tissue. 58,85 The NIR<br />

dye indocyanine green (ICG), a model diagnostic agent, has been conjugated with the side<br />

carboxylic acid of branched PG having a PAMAM core and the terminal folic acid group as a<br />

targeting moiety (Figure 10.3). 86 The resulting conjugate, PAMAM16–PG–(ICG)–folate, exhibited<br />

selective binding to KB cells (overexpressing folate receptors) but not to SK-Br3 cells that cannot<br />

express folate receptors. Additionally, this selective binding was partially blocked <strong>by</strong> free<br />

folic acid.<br />

10.7 CONCLUSIONS<br />

HN<br />

O<br />

NH<br />

(CH2 ) 2<br />

m<br />

O<br />

N<br />

H<br />

O<br />

n<br />

NH OC (CH2 ) 2<br />

COOH<br />

H<br />

N<br />

O<br />

+<br />

N<br />

Br –<br />

N<br />

CO<br />

HN<br />

SO 3 –<br />

COOH<br />

COOH<br />

Owing to its favorable physicochemical properties, PG has been shown to be an excellent polymeric<br />

carrier of both diagnostic and therapeutic agents. PG–TXL has become the first PG-based<br />

polymeric agent to be tested in clinical trials. The newer generation of PG-based polymers will have<br />

N<br />

N<br />

H 2 N<br />

HN<br />

N<br />

N<br />

Folic acid<br />

FIGURE 10.3 Structure of PAMAM16–PG–(ICG)–folate containing folic acid molecules at the termini of the<br />

polymer and dye molecules (ICG-NH 2) at the side chains of the polymer. (From Tansey, W., Ke, S., Cao, X. Y.<br />

et al., J. Control. Rel., 94, 39–51, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


196<br />

to meet a number of challenges, including the development of novel polymers with modulated rates<br />

of degradation, versatile conjugation chemistry allowing site-specific attachment of targeting<br />

moieties, and polymerization methods that allow accurate control of polymer molecular weights<br />

and molecular weight distributions.<br />

ACKNOWLEDGMENTS<br />

This work was supported in part <strong>by</strong> the National Cancer Institute (R29 CA74819), the Department<br />

of Defense (BC960384), and the John S. Dunn Foundation. We thank Donald R. Norwood for<br />

editing the manuscript.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Li, C., Poly(L-glutamic acid)-anticancer drug conjugates, Adv. Drug Deliv. Rev., 54, 695–713, 2002.<br />

2. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2, 347–360, 2003.<br />

3. Allen, T. M. and Cullis, P. R., Drug delivery systems: Entering the mainstream, Science, 303,<br />

1818–1822, 2004.<br />

4. Idelson, M. E. and Blout, E. R., High molecular weight poly(L-glutamic acid): Preparation and<br />

physical rotation changes, J. Am. Chem. Soc., 80, 4631–4634, 1958.<br />

5. Fuller, W. D., Verlander, M. S., and Goodman, M., A procedure for the facile synthesis of amino-acid<br />

N-carboxyanhydrides, Biopolymers, 15, 1869–1871, 1976.<br />

6. Katakai, R. and Lisuka, Y., An improved method for the synthesis of N-carboxy amino acid<br />

anhydrides using trichloromethyl chloroformate, J. Org. Chem., 50, 715–716, 1985.<br />

7. Block, H., Poly(g-benzyl-L-glutamate) and Other Glutamic Acid Containing Polymers, Gordon and<br />

Breach, New York, 1983.<br />

8. Deming, T. J., Facile synthesis of block copolypeptides of defined architecture, Nature, 390, 386–389,<br />

1997.<br />

9. Poche, D., Moore, M., and Bowles, J., An unconventional method for purifying the N-carboxyanhydride<br />

derivatives of g-alkyl-L-glutamates, Synth. Commun., 29, 843–854, 1999.<br />

10. Aliferis, T., Iatrou, H., and Hadjichristidis, N., Living polypeptides, Biomacromolecules, 5,<br />

1653–1656, 2004.<br />

11. Han<strong>by</strong>, W. E., Waley, S. G., and Watson, J., Synthetic polypeptides. Part II. Polyglutamic acid,<br />

J. Chem. Soc., 3239–3249, 1950.<br />

12. Palacios, P., Bussat, P., and Bichon, D., Novel solid-phase synthesis of thiol-terminated-poly(alphaamino<br />

acid)–drug conjugate, J. Biochem. Biophys. Methods, 23, 67–72, 1991.<br />

13. Vega, J., Ke, S., Fan, Z. et al., Targeting doxorubicin to epidermal growth factor receptors <strong>by</strong> sitespecific<br />

conjugation of C225 to poly(L-glutamic acid) through a polyethylene glycol spacer, Pharm.<br />

Res., 20, 826–832, 2003.<br />

14. Nishiyama, N., Okazaki, S., Cabral, H. et al., Novel cisplatin-incorporated polymeric micelles can<br />

eradicate solid tumors in mice, Cancer Res., 63, 8977–8983, 2003.<br />

15. De Marre, A. and Schacht, E., Preparation of 4-nitrophenol carborate esters of poly[N-5-(2-hydroxyethyl)-L-glutamine]<br />

and coupling with bioactive agents, Markromol. Chem., 193, 3023–3030,<br />

1992.<br />

16. Romberg, B., Metselaar, J. M., deVringer, T. et al., Enzymatic degradation of liposome-grafted<br />

poly(hydroxyethyl L-glutamine), Bioconjug. Chem., 16, 767–774, 2005.<br />

17. D’Souza, A. J. and Topp, E. M., Release from polymeric prodrugs: Linkages and their degradation,<br />

J. Pharm. Sci., 93, 1962–1979, 2004.<br />

18. Thomssen, C., Schmit, M., Goretzki, L. et al., Prognostic value of the cysteine proteases cathepsins B<br />

and cathepsin L in human breast cancer, Clin. Cancer Res., 1, 741–746, 1995.<br />

19. Mai, J., Waisman, D. M., and Sloane, B. F., Cell surface complex of cathepsin B/annexin II tetramer in<br />

malignant progression, Biochim. Biophys. Acta, 1477, 215–230, 2000.<br />

20. Kornguth, S. E., Kalinke, T., Robins, H. I., Cohen, J. D., and Turski, P., Preferential binding of<br />

radiolabeled poly-L-lysines to C6 and U87 MG glioblastomas compared with endothelial cells<br />

in vitro, Cancer Res., 49, 6390–6395, 1989.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 197<br />

21. Hoste, K., De Winne, K., and Schacht, E., Polymeric prodrugs, Int. J. Pharm., 277, 119–131, 2004.<br />

22. Chiu, H. C., Kopeckova, P., Deshmane, S. S., and Kopecek, J., Lysosomal degradability of poly(alphaamino<br />

acids), J. Biomed. Mater. Res., 34, 381–392, 1997.<br />

23. Tsutsumi, A., Perly, B., Forchioni, A., and Charahaty, C., A magnetic resonance study of the<br />

segmental motion and local conformations of poly(L-glutamic acid) in aqueous solutions, Macromolecules,<br />

11, 977–986, 1978.<br />

24. Dolnik, V., Novotny, M., and Chmelik, J., Electromigration behavior of poly(L-glutamate) conformers<br />

in concentrated polyacrylamide gels, Biopolymers, 33, 1299–1306, 1993.<br />

25. Miller, W. G., Degradation of synthetic polypeptides II. degradation of poly-L-glutamic acid <strong>by</strong><br />

proteolytic enzymes in 0.20 M sodium chloride, J. Am. Chem. Soc., 86, 3913–3918, 1964.<br />

26. Kishore, B. K., Kallay, Z., Lambricht, P., Laurent, G., and Tulkens, P. M., Mechanism of protection<br />

afforded <strong>by</strong> polyaspartic acid against gentamicin-induced phospholipidosis I. Polyaspartic acid binds<br />

gentamicin and displaces it from negatively charged phospholipid layers in vitro, J. Pharmacol. Exp.<br />

Ther., 255, 867–874, 1990.<br />

27. McCormick-Thomson, L. A. and Duncan, R., Poly(amino acid) copolymers as a potential soluble drug<br />

delivery system. I: Pinocytotic uptake and lysosomal degradation measured in vitro, J. Bioact.<br />

Compat. Polym., 4, 242–251, 1989.<br />

28. Hoes, C. J. T., Potman, W., Van Heeswijk, W. A. R. et al., Optimization of macromolecular prodrugs<br />

of the antitumor antibiotic Adriamycin, J. Control. Rel., 2, 205–213, 1985.<br />

29. Singer, J. W., Nudelman, E., De Vries, P., Shaffer, S., Metabolism of poly-L-glutamic acid (PG)<br />

paclitaxel (CT-2103): Intratumoral peptide cleavage followed <strong>by</strong> esterolysis, In 5th International<br />

Symposium on Polymer Therapeutics: From Laboratory to Clinical Practice, Cardiff, UK, 2002.<br />

30. Van Heeswijk, W. A. R., Hoes, C. J. T., Stoffer, T. et al., The synthesis and characterization of<br />

polypeptide–adriamycin conjugates and its complexes with Adriamycin. Part 1, J. Control. Rel., 1,<br />

301–315, 1985.<br />

31. Hoes, C. J. T., Grootoonk, J., Duncan, R. et al., Biological properties of adriamycin bound to<br />

biodegradable polymeric carriers, J. Control. Rel., 23, 37–54, 1993.<br />

32. Zunino, F., Pratesi, G., and Micheloni, A., Poly(carboxylic acid) polymers as carriers for anthracyclines,<br />

J. Control. Rel., 10, 65–73, 1989.<br />

33. Hurwitz, E., Wilchek, M., and Pitha, J., Soluble macromolecules as carriers for daunorubicin, J. Appl.<br />

Biochem., 2, 25–35, 1980.<br />

34. Kato, Y., Saito, M., Fukushima, H., Takeda, Y., and Hara, T., Antitumor activity of 1-beta-D-arabinofuranosylcytosine<br />

conjugated with polyglutamic acid and its derivative, Cancer Res., 44, 25–30,<br />

1984.<br />

35. Batz, H. G., Ringsdorf, H., and Ritter, H., Pharmacologically active polymers. 7. Cyclophosphamideand<br />

steroid hormone-containing polymers as potential anticancer compounds, Makromol. Chem., 175,<br />

2229–2239, 1974.<br />

36. Moromoto, Y., Sugibayashi, K., Sugihara, S. et al., Antitumor agent poly(amino acid) conjugates as a<br />

drug carrier in chemotherapy, J. Pharm. Dyn., 7, 688–698, 1984.<br />

37. Roos, C. F., Matsumoto, S., Takarura, Y., Hashida, M., and Sezaki, H., Physicochemical and antitumor<br />

characteristics of some polyamino acid prodrugs of mitomycin C, Int. J. Pharm., 22, 75–87,<br />

1984.<br />

38. Soyez, H., Schacht, E., and Vanderkerken, S., The crucial role of spacer groups in macromolecular<br />

prodrug design, Adv. Drug Deliv. Rev., 21, 81–106, 1996.<br />

39. de Marre, A., Seymour, L. W., and Schacht, E., Evaluation of the hydrolytic and enzymatic stability of<br />

macromolecular mitomycin C derivatives, J. Control. Rel., 31, 89–97, 1994.<br />

40. Avichezer, D., Schechter, B., and Arnon, R., Functional polymers in drug delivery: Carrier-supported<br />

CDDP (cisplatin) complexes of polycarboxylates-effects on human ovarian carcinoma, React. Funct.<br />

Polym., 36, 59–69, 1998.<br />

41. Schlechter, B., Neumann, A., Wilchek, M., and Arnon, R., Soluble polymers as carriers of cisplatinum,<br />

J. Control. Rel., 10, 75–87, 1989.<br />

42. Seymour, L. W., Soyez, H., De Marre, A., Shoaibi, M. A., and Schacht, E. H., Polymeric prodrugs of<br />

mitomycin C designed for tumour tropism and sustained activation, Anti-cancer Drug Des., 11,<br />

351–365, 1996.<br />

43. Horwitz, S. B., Taxol (Paclitaxel): Mechanisms of action, Ann. Oncol., 5, S3–S6, 1994.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


198<br />

Nanotechnology for Cancer Therapy<br />

44. Li, C., Yu, D. F., Newman, R. A. et al., Complete regression of well-established tumors using a novel<br />

water-soluble poly(L-glutamic acid)–paclitaxel conjugate, Cancer Res., 58, 2404–2409, 1998.<br />

45. Li, C., Newman, R. A., Wu, Q. P. et al., Biodistribution of paclitaxel and poly(L-glutamic acid)–<br />

paclitaxel conjugate in mice with ovarian OCa-1 tumor, Cancer Chemother. Pharmacol., 46,<br />

416–422, 2000.<br />

46. Oldham, E. A., Li, C., Ke, S., Wallace, S., and Huang, P., Comparison of action of paclitaxel and poly(Lglutamic<br />

acid)–paclitaxel conjugate in human breast cancer cells, Int. J. Oncol., 16, 125–132, 2000.<br />

47. Singer, J. W., Shaffer, S., Baker, B. et al., Paclitaxel poliglumex (XYOTAX; CT-2103): An intracellularly<br />

targeted taxane, Anticancer Drugs, 16, 243–254, 2005.<br />

48. Li, C., Price, J. E., Milas, L. et al., Antitumor activity of poly(L–glutamic acid)–paclitaxel on<br />

syngeneic and xenografted tumors, Clin. Cancer Res., 5, 891–897, 1999.<br />

49. Auzenne, E., Donato, N. J., Li, C. et al., Superior therapeutic profile of poly-L-glutamic acid–paclitaxel<br />

copolymer compared with taxol in xenogeneic compartmental models of human ovarian<br />

carcinoma, Clin. Cancer Res., 8, 573–581, 2002.<br />

50. Todd, R., Sludden, J., Boddy, A. V. et al., Phase I and pharmacological study of CT-2103, a poly(Lglutamic<br />

acid)–paclitaxel conjugate, Proc. Am. Soc. Clin. Oncol., 20, 2001 (abstr 439).<br />

51. Bernareggi, A., Oldham, F., Barone, C., Tumor-targeted taxane: PK evidence for an enhanced permeability<br />

and retention effect with paclitaxel poliglumex (CT-2103), Presented at the European<br />

Society for Medical Oncology (ESMO), October 29–November 2, 2004, Vienna, Austria.<br />

52. Markman, M., Improving the toxicity profile of chemotherapy for advanced ovarian cancer: A<br />

potential role for CT-2103, J. Exp. Ther. Oncol., 4, 131–136, 2004.<br />

53. Sabbatini, P., Aghajanian, C., Dizon, D. et al., Phase II study of CT-2103 in patients with recurrent<br />

epithelial ovarian, fallopian tube, or primary peritoneal carcinoma, J. Clin. Oncol., 22, 4523–4531,<br />

2004.<br />

54. Zou, Y., Wu, Q. P., Tansey, W. et al., Effectiveness of water soluble poly(L-glutamic acid)–camptothecin<br />

conjugate against resistant human lung cancer xenografted in nude mice, Int. J. Oncol., 18,<br />

331–336, 2001.<br />

55. Singer, J. W., De Vries, P., Bhatt, R. et al., Conjugation of camptothecins to poly(L-glutamic acid),<br />

Ann. N Y Acad. Sci., 922, 136–150, 2000.<br />

56. Singer, J. W., Bhatt, R., Tulinsky, J. et al., Water-soluble poly(L-glutamic acid)-Gly-camptothecin<br />

conjugates enhance camptothecin stability and efficacy in vivo, J. Control. Rel., 74, 243–247, 2001.<br />

57. Bhatt, R., de Vries, P., Tulinsky, J. et al., Synthesis and in vivo antitumor activity of poly(L-glutamic<br />

acid) conjugates of 20S-camptothecin, J. Med. Chem., 46, 190–193, 2003.<br />

58. Hawrysz, D. J. and Sevick-Muraca, E. M., Developments toward diagnostic breast cancer imaging<br />

using near-infrared optical measurements and fluorescent contrast agents, Neoplasia, 2, 388–417,<br />

2000.<br />

59. Li, C., Ke, S., Wu, Q. P. et al., Tumor irradiation enhances the tumor-specific distribution of<br />

poly(L-glutamic acid)–conjugated paclitaxel and its antitumor efficacy, Clin. Cancer Res., 6,<br />

2829–2834, 2000.<br />

60. Ke, S., Milas, L., Charnsangavej, C., Wallace, S., and Li, C., Potentiation of radioresponse <strong>by</strong><br />

polymer–drug conjugates, J. Control. Rel., 74, 237–242, 2001.<br />

61. Milas, L., Mason, K. A., Hunter, N., Li, C., and Wallace, S., Poly(L-glutamic acid)–paclitaxel conjugate<br />

is a potent enhancer of tumor radiocurability, Int. J. Radiat. Oncol. Biol. Phys., 55, 707–712,<br />

2003.<br />

62. Tishler, R. B., Polymer–conjugated paclitaxel as a radiosensitizing agent-a big step forward for<br />

combined modality therapy?, Int. J. Radiat. Oncol. Biol. Phys., 55, 563–564, 2003.<br />

63. Bogdanov, A. A., Lewin, M., and Weissleder, R., Approaches and agents for imaging the vascular<br />

system, Adv. Drug Deliv. Rev., 37, 279–293, 1993.<br />

64. Kroft, L. J. and de Roos, A., Blood pool contrast agents for cardiovascular MR imaging, J. Magn.<br />

Reson. Imaging, 10, 395–403, 1999.<br />

65. Brasch, R. C., New directions in the development of MR imaging contrast media, Radiology, 183,<br />

1–11, 1992.<br />

66. Su, M. Y., Muhler, A., Lao, X., and Nalcioglu, O., Tumor characterization with dynamic contrastenhanced<br />

MRI using MR contrast agents of various molecular weights, Magn. Reson. Med., 39,<br />

259–269, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(L-Glutamic Acid): Efficient Carrier of Cancer Therapeutics and Diagnostics 199<br />

67. van Dijke, C. F., Brasch, R. C., Roberts, T. P. et al., Mammary carcinoma model: Correlation of<br />

macromolecular contrast-enhanced MR imaging characterizations of tumor microvasculature and<br />

histologic capillary density, Radiology, 198, 813–818, 1996.<br />

68. Gossmann, A., Okuhata, Y., Shames, D. M. et al., Prostate cancer tumor grade differentiation with<br />

dynamic contrast-enhanced MR imaging in the rat: Comparison of macromolecular and small-molecular<br />

contrast media-preliminary experience, Radiology, 213, 265–272, 1999.<br />

69. Daldrup, H., Shames, D. M., Wendland, M. et al., Correlation of dynamic contrast-enhanced magnetic<br />

resonance imaging with histologic tumor grade: Comparison of macromolecular and small-molecular<br />

contrast media, Pediatr. Radiol., 28, 67–78, 1998.<br />

70. van Bemmel, C. M., Wink, O., Verdonck, B., Viergever, M. A., and Niessen, W. J., Blood pool<br />

contrast-enhanced MRA: Improved arterial visualization in the steady state, IEEE Trans. Med.<br />

Imaging, 22, 645–652, 2003.<br />

71. Schuhmann-Giampieri, G., Schmitt-Willich, H., Frenzel, T., Press, W. R., and Weinmann, H. J.,<br />

In vivo and in vitro evaluation of Gd–DTPA–polylysine as a macromolecular contrast agent for<br />

magnetic resonance imaging, Invest. Radiol., 26, 969–974, 1991.<br />

72. Bogdanov, J., Alexei, A., Weissleder, R., and Brady, T. J., Long-circulating blood pool imaging<br />

agents, Adv. Drug Deliv. Rev., 16, 335–348, 1995.<br />

73. Wang, S. C., Wikstrom, M. G., White, D. L. et al., Evaluation of Gd–DTPA–labeled dextran as an<br />

intravascular MR contrast agent: Imaging characteristics in normal rat tissues, Radiology, 175,<br />

483–488, 1990.<br />

74. Rebizak, R., Schaefer, M., and Dellacherie, E., Polymeric conjugates of Gd(3C)-diethylenetriaminepentaacetic<br />

acid and dextran. 1. Synthesis, characterization, and paramagnetic properties,<br />

Bioconjug. Chem., 8, 605–610, 1997.<br />

75. Wiener, E. C., Brechbiel, M. W., Brothers, H. et al., Dendrimer-based metal chelates: A new class of<br />

magnetic resonance imaging contrast agents, Magn. Reson. Med., 31, 1–8, 1994.<br />

76. Adam, G., Neuerburg, J., Spuntrup, E. et al., Gd–DTPA–cascade-polymer: Potential blood pool<br />

contrast agent for MR imaging, J. Magn. Reson. Imaging, 4, 462–466, 1994.<br />

77. Kobayashi, H. M. and Brechbiel, W., Dendrimer-based macromolecular MRI contrast agents: Characteristics<br />

and application, Mol. Imaging, 2, 1–10, 2003.<br />

78. Kim, Y. H., Choi, B. I., Cho, W. H. et al., Dynamic contrast-enhanced MR imaging of VX2 carcinomas<br />

after X-irradiation in rabbits: Comparison of gadopentetate dimeglumine and a<br />

macromolecular contrast agent, Invest. Radiol., 38, 539–549, 2003.<br />

79. Unger, E. C., Shen, D., Wu, G. et al., Gadolinium-containing copolymeric chelates—a new potential<br />

MR contrast agent, Magma, 8, 154–162, 1999.<br />

80. Gupta, H., Wilkinson, R. A., Bogdanov, A. A., Callahan, R. J., and Weissleder, R., Inflammation:<br />

Imaging with methoxy poly(ethylene glycol)–poly-L-lysine–DTPA, a long-circulating graft copolymer,<br />

Radiology, 197, 665–669, 1995.<br />

81. Rebizak, R., Schaefer, M., and Dellacherie, E., Polymeric conjugates of Gd(3C)-diethylenetriaminepentaacetic<br />

acid and dextran. 2. Influence of spacer arm length and conjugate molecular mass on the<br />

paramagnetic properties and some biological parameters, Bioconjug. Chem., 9, 94–99, 1998.<br />

82. Franano, F. N., Edwards, W. B., Welch, M. J. et al., Biodistribution and metabolism of targeted and<br />

nontargeted protein-chelate–gadolinium complexes: Evidence for gadolinium dissociation in vitro<br />

and in vivo, Magn. Reson. Imaging, 13, 201–214, 1995.<br />

83. Wen, X., Jackson, E. F., Price, R. E. et al., Synthesis and characterization of poly(L-glutamic acid)<br />

gadolinium chelate: A new biodegradable MRI contrast agent, Bioconjug. Chem., 15, 1408–1415, 2004.<br />

84. Lu, Z. R., Wang, X., Parker, D. L., Goodrich, K. C., and Buswell, H. R., Poly(L-glutamic acid)<br />

Gd(III)–DOTA conjugate with a degradable spacer for magnetic resonance imaging, Bioconjug.<br />

Chem., 14, 715–719, 2003.<br />

85. Ke, S., Wen, X., Gurfinkel, M. et al., Near-infrared optical imaging of epidermal growth factor<br />

receptor in breast cancer xenografts, Cancer Res., 63, 7870–7875, 2003.<br />

86. Tansey, W., Ke, S., Cao, X. Y. et al., Synthesis and characterization of branched poly(L-glutamic acid)<br />

as a biodegradable drug carrier, J. Control. Rel., 94, 39–51, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


11<br />

CONTENTS<br />

Noninvasive Visualization of<br />

In Vivo Drug Delivery of<br />

Paramagnetic Polymer<br />

Conjugates with MRI<br />

Zheng-Rong Lu, Yanli Wang, Furong Ye, Anagha<br />

Vaidya, and Eun-Kee Jeong<br />

11.1 Introduction ....................................................................................................................... 201<br />

11.2 Magnetic Resonance Imaging ........................................................................................... 202<br />

11.2.1 Principles of MRI ................................................................................................ 202<br />

11.2.2 Contrast-Enhanced MRI...................................................................................... 203<br />

11.3 MR Imaging of Paramagnetic HPMA Copolymer Conjugates........................................ 204<br />

11.3.1 MR Imaging of a Paramagnetic HPMA Conjugate in an Animal<br />

Tumor Model....................................................................................................... 204<br />

11.3.2 MR Imaging of a Paramagnetic HPMA Conjugate for Bone Delivery............. 206<br />

11.4 MR Imaging of Paramagnetic Poly(L-Glutamic Acid) Conjugates ................................. 206<br />

11.4.1 MR Imaging of PGA–(Gd-DOTA) Conjugate in an Animal<br />

Tumor Model....................................................................................................... 206<br />

11.4.2 MR Imaging of PGA–Mce6–(Gd-DOTA) Conjugate in an Animal<br />

Tumor Model....................................................................................................... 208<br />

11.5 Summary............................................................................................................................ 210<br />

References .................................................................................................................................... 210<br />

11.1 INTRODUCTION<br />

The conjugation of therapeutics to water-soluble biomedical polymers increases the aqueous solubility<br />

of hydrophobic drugs, prolongs in vivo drug retention time, reduces systemic toxicity, and<br />

enhances therapeutic efficacy. 1,2 Polymer drug conjugates can preferentially accumulate in solid<br />

tumor tissues due to the hyperpermeability of tumor blood vessels or the so-called “enhanced<br />

permeability and retention (EPR) effect.” 3,4 Polymer drug conjugates can also down-regulate or<br />

overcome multiple drug resistance. 5,6 Because of these unique properties, polymer drug conjugates<br />

exhibit higher therapeutic efficacy than the corresponding therapeutics alone. Several polymer drug<br />

conjugates are currently used in clinical cancer treatment, and more are in the pipeline of clinical<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

201


202<br />

development. For example, styrene–maleic acid copolymer neocarzinostatin conjugate (SMANCS)<br />

is used for liver cancer treatment in Japan; 7 pegylated adenosine deaminase 8 and asparaginase 9,10<br />

are used for enzyme replacement therapy for immunodeficiency and for the treatment of acute<br />

lymphoblastic leukemia, respectively. Many other polymer drug conjugates, including pegylated<br />

interferon a, 11 PEG–camptothecin conjugate, 12 poly(L-glutamic acid) paclitaxel conjugate, 13<br />

poly[N-(2-hydroxypropyl)methacrylamide] (PHPMA)-cis-platinate conjugate, 14 and PHPMA–<br />

doxorubicin conjugate 15,16 are in various phases of clinical trials.<br />

The in vivo behavior, including pharmacokinetics, biodistribution, drug delivery efficiency and<br />

therapeutic efficacy, of polymer drug conjugates has been traditionally evaluated <strong>by</strong> using blood<br />

and urine sampling, biopsy-based methods, and symptom-based observations. These methods are<br />

sometimes invasive and cannot accurately provide real time information on the interaction of<br />

polymers with various organs and tissues, the targeting and delivery efficiency of the drug delivery<br />

system, and therapeutic response. A large number of animals are also required in preclinical<br />

development. Sometimes, the data obtained <strong>by</strong> conventional biopsies may be misleading, which<br />

might be one of the causes of the efficacy discrepancy between preclinical development in animal<br />

models and human clinical trials and the failure of some clinical trials due to improper selection of<br />

drug candidates.<br />

In vivo drug delivery <strong>by</strong> polymer drug conjugates involves circulation of the conjugates in the<br />

blood; interaction with the major organs including the liver, heart, lungs, spleen and kidneys, etc.;<br />

transport to target tissue; and uptake <strong>by</strong> target cells and drug release. Direct and continuous<br />

evaluation of drug delivery efficiency and tissue interaction of polymeric drug conjugates is critical<br />

to develop more efficacious drug delivery systems. Recent advancements in biomedical imaging<br />

technology have provided the essential tools for noninvasive and continuous in vivo evaluation for<br />

drug delivery. Nuclear medicine, including positron emission tomography (PET) and single photon<br />

emission computed tomography (SPECT), is a clinical imaging modality with high detection<br />

sensitivity (ca. 10 K9 M). Magnetic resonance imaging (MRI) is a noninvasive clinical imaging<br />

modality with high spatial resolution. Both imaging modalities are able to noninvasively visualize<br />

in vivo drug delivery with polymeric drug conjugates. The polymer conjugates can be labeled with<br />

imaging probes and noninvasively and continuously monitored with the imaging modalities.<br />

The number of experimental animals can be dramatically reduced in the preclinical development<br />

of the conjugates. Biomedical imaging will provide real-time information of the in vivo behavior of<br />

polymeric drug conjugates. In this chapter, we will focus on noninvasive visualization of in vivo<br />

drug delivery with paramagnetic polymeric conjugates and contrast-enhanced MRI.<br />

11.2 MAGNETIC RESONANCE IMAGING<br />

11.2.1 PRINCIPLES OF MRI<br />

Nanotechnology for Cancer Therapy<br />

MRI is a clinical imaging modality that produces three-dimensional anatomic images with high<br />

spatial resolution and provides in vivo physiological properties including physiochemical information,<br />

flow diffusion, and motion in the tissues. The physics of MRI is complicated and has been<br />

detailed in a number of monographs. 17,18 The fundamental principles of MRI are briefly reviewed<br />

in this section. A normal human adult has approximately 60% of the body weight as water. Water<br />

protons have a magnetic moment and the orientations of the magnetic moments are random in the<br />

absence of an external magnetic field. When a patient is placed in a MRI scanner, the proton<br />

magnetic moments align either along or against the static magnetic field (B0) of the scanner and<br />

create a net magnetization pointing in the direction of the main magnetic field of the scanner.<br />

The magnitude of the magnetization is proportional to the external magnetic field strength as well<br />

as the amount of protons.<br />

Nuclear magnetic resonance (NMR) measures the change of the magnetization <strong>by</strong> applying<br />

radiofrequency (RF) pulses. When an RF pulse is applied to create an oscillating electromagnetic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Noninvasive Visualization of In Vivo Drug Delivery of Paramagnetic Polymer Conjugates 203<br />

field (B1) perpendicular to the main field, the longitudinal magnetization is tipped away from the<br />

static magnetic field and deviates from the equilibrium state (in longitudinal space). A perfect 908<br />

RF pulse flips the total longitudinal magnetization onto the transverse plane. Immediately after<br />

excitation with the RF pulse, the transverse magnetization then undergoes the decay process.<br />

The decaying transverse magnetization induces an NMR signal as an electromotive force (emf)<br />

on an RF coil. Simultaneously, the nuclear spin system approaches toward the equilibrium state<br />

with a characteristic recovery time T1, which depends upon the physical and chemical environment<br />

of the molecules.<br />

Longitudinal T1 relaxation involves the return of protons from the high energy state back to<br />

equilibrium <strong>by</strong> dissipating their excess energy to their surroundings. The process is also called spinlattice<br />

relaxation. Transverse (T 2) relaxation involves the transfer of the spin angular momentum<br />

among the protons via the interactions such as dipole–dipole interaction. The process is called<br />

spin–spin relaxation. In biological systems, the T2 relaxation time is typically shorter than T1<br />

relaxation time. Because the main magnetic field of MRI scanners is not perfectly homogenous,<br />

the inhomogeneity of the magnetic field also causes dephasing of individual magnetizations of<br />

protons, resulting in more rapid loss of transverse magnetization. The process is called<br />

T 2 relaxation.<br />

In MR imaging, the position dependent frequency and phase are encoded into the transverse<br />

magnetization, and the measured signals are Fourier transformed to construct the spatial map<br />

(image). MR images are created based on the proton density, or T 1 and T 2 relaxation rates, and<br />

flow and diffusion properties in different tissues. These parameters vary between different tissues<br />

and create image contrast among the tissues. Tissues with a short T1 relaxation time give a strong<br />

MR signal, resulting in bright images in T1-weighted MRI. Tissues with a short T2 relaxation time<br />

produce a weak MR signal, resulting in dark images in T2-weighted MRI.<br />

11.2.2 CONTRAST-ENHANCED MRI<br />

In many cases, the differences between the MR signal intensities of normal and diseased tissues are<br />

not large enough to produce obvious contrast for definitive diagnosis. Contrast agents are used to<br />

enhance the image contrast between normal and diseased tissues to improve diagnostic sensitivity<br />

and specificity. 19,20 MRI contrast agents are either paramagnetic metal chelates or ultrasmall<br />

superparamagnetic iron oxides. These agents interact with surrounding water molecules and<br />

increase the relaxation rates (1/T1 and 1/T2) of water protons. The change of relaxation rate is<br />

linearly proportional to the concentration of contrast agents, [M]:<br />

1=T i;obs Z 1=T i;d Cr i½MŠ; (11.1)<br />

where iZ1,2. In Equation 11.1, 1/T i,obs is the observed relaxation rate with a contrast agent, 1/T i,d<br />

the relaxation rate without the contrast agent, and ri the relaxivity of the contrast agent. Relaxivity is<br />

a measurement of the efficiency of a contrast agent to enhance the relaxation rate of water protons<br />

or the efficiency to generate image contrast enhancement. Because the uptake of contrast agents<br />

varies between different tissues, MR image contrast is enhanced in the tissue with a high contrast<br />

agent concentration compared to that with a low concentration.<br />

MRI contrast agents approved for clinical applications are mainly stable paramagnetic chelates,<br />

e.g., Gd(III) chelates 21–24 and Mn(II) chelates, 25 and ultrasmall superparamagnetic iron oxide<br />

(USPIO). 26 Gd(III) chelates of DTPA, DOTA, or their derivatives are commonly used for<br />

contrast-enhanced MRI. Both Gd(III) and Mn(II) chelates are mainly used as T1 contrast agents,<br />

whereas USPIO is used as a T2 contrast agent. These agents significantly enhance image contrast<br />

between normal tissue and diseased tissue and improve the diagnostic accuracy. Currently, contrast<br />

agents are used in approximately 30% of clinical MRI examinations. Macromolecular Gd(III)<br />

complexes have also been developed as MRI contrast agents. 27,28 These agents have a prolonged<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


204<br />

blood circulation time and preferentially accumulate in solid tumors due to the EPR effect. Macromolecular<br />

Gd(III) complexes have demonstrated superior contrast enhancement in MR<br />

angiography and cancer imaging in animal models, but are not yet approved for clinical<br />

applications.<br />

Contrast-enhanced MRI is a useful method for noninvasive in vivo evaluation of polymeric<br />

drug delivery systems after the systems are labeled with MRI contrast agents. The real-time<br />

pharmacokinetics and biodistribution of the labeled drug delivery systems can be continuously<br />

visualized <strong>by</strong> contrast-enhanced MRI. It has a potential to provide accurate four-dimensional<br />

information of in vivo properties of the drug delivery systems.<br />

11.3 MR IMAGING OF PARAMAGNETIC HPMA COPOLYMER CONJUGATES<br />

N-(2-Hydroxypropyl)methacrylamide (HPMA) copolymer is a biocompatible water-soluble polymeric<br />

carrier for therapeutics. 1 The conjugation of anti-cancer drugs to HPMA copolymers results<br />

in many advantageous features over small molecular therapeutics, including improved water solubility<br />

and bioavailability, preferential accumulation of the conjugates in solid tumors, reduced<br />

nonspecific toxicity, and down-regulation of multi-drug resistance. 1,29 A number of HPMA copolymer<br />

drug conjugates have been systematically investigated both in vitro and in vivo. Several<br />

HPMA copolymer-anti-cancer drug conjugates are now in various phases of clinical development<br />

for cancer treatment. 30 The incorporation of MRI contrast agents into HPMA copolymer conjugates<br />

allows direct visualization of real-time drug delivery of the conjugates in animal models with<br />

contrast-enhanced MRI. The noninvasive imaging reveals some interesting features of drug<br />

delivery with the conjugates that cannot be obtained with conventional surgery based evaluations.<br />

11.3.1 MR IMAGING OF A PARAMAGNETIC HPMA CONJUGATE<br />

IN AN ANIMAL TUMOR MODEL<br />

Nanotechnology for Cancer Therapy<br />

HPMA copolymers have been labeled with radioactive probes including 131 Iand 99m Tc for<br />

noninvasive visualization of in vivo drug delivery of the conjugates with g-scintigraphy. 31–33<br />

Nuclear medicine is highly sensitive for detecting labeled conjugates, but its poor spatial resolution<br />

cannot provide detailed biodistribution of conjugates in tissues and organs. In comparison, contrastenhanced<br />

MRI provides clear visualization of the biodistribution of paramagnetically labeled<br />

conjugates with high spatial resolution.<br />

HPMA copolymers were labeled with an MRI contrast agent, Gd-DOTA, and investigated in<br />

an animal tumor model with contrast-enhanced MRI. 34 The structure of a labeled conjugate,<br />

poly[HPMA-co-(MA-Gly-Gly-1,6-hexanediamine-(Gd-DOTA))], is shown in Figure 11.1.<br />

The molecular weight of the polymer conjugate was 25.6 kDa (PDZ1.50) and the Gd content<br />

was 0.33 mmol Gd/g polymer. The T1 relaxivity of the conjugate was 6.0 mM K1 s K1 per attached<br />

Gd(III) chelate at 3 T. Figure 11.2 shows the dynamic, contrast-enhanced, three-dimensional<br />

maximum intensity projection (MIP) (Figure 11.2a) and 2D coronal MR images (Figure 11.2b)<br />

of a mouse bearing a human prostate carcinoma DU-145 xenograft and a Kaposi’s sarcoma xenograft.<br />

The conjugate was administrated intravenously at a Gd equivalent dose of 0.1 mmo1/kg body<br />

weight. The pharmacokinetics and biodistribution of the conjugate were clearly revealed in the<br />

contrast-enhanced MR images.<br />

Strong MRI signal was observed in the heart, liver, kidneys and vasculature in the 3D images at<br />

the initial stage after the injection of the conjugates. The signal intensity decreased over a 60-min<br />

period; meanwhile, the signal intensity in the urinary bladder increased gradually, indicating that<br />

the conjugate was excreted via renal filtration. Significant contrast enhancement was still visible in<br />

the heart at 60 min. After 22 h, most of the conjugate was cleared from the body except the liver, as<br />

shown in Figure 11.2. The signal intensity in the liver was stronger at 22 h than in the precontrast<br />

image, suggesting significant interaction of HPMA copolymers with the liver.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Noninvasive Visualization of In Vivo Drug Delivery of Paramagnetic Polymer Conjugates 205<br />

CH2<br />

CH3<br />

C<br />

m<br />

O<br />

NH<br />

CH2<br />

CH OH<br />

CH3<br />

CH 2<br />

CH3<br />

C<br />

n<br />

O<br />

NH<br />

CH2<br />

C O<br />

NH<br />

CH2<br />

O O<br />

NH(CH2) 6NH C<br />

− OOC<br />

N N<br />

Gd<br />

N N<br />

3+<br />

COO −<br />

COO −<br />

FIGURE 11.1 The structure of poly[HPMA-co-(MA-Gly-Gly-1,6-hexanediamine-(Gd-DOTA))].<br />

The coronal images cross-sectioning the tumor tissues showed heterogeneous uptake of HPMA<br />

copolymers in two different tumors (Figure 11.2b). The prostate carcinoma had a smaller size, but<br />

stronger MR signal than the Kaposi’s sarcoma. For Kaposi’s sarcoma, the contrast enhancement<br />

was mainly observed in the periphery of the tumor tissue and to a lower extent in the inner tumor<br />

tissue. A plausible explanation is that the interstitium of the large tumor was possibly necrotic,<br />

which would limit the access of the conjugate. The results suggested that tumor uptake of HPMA<br />

copolymers may vary from tumor to tumor and with different tumor sizes. It might affect the<br />

efficacy of the polymer conjugates and more detailed studies are needed to understand the<br />

possible correlations.<br />

FIGURE 11.2 Three-dimensional maximum intensity projection (MIP) (a) and 2D coronal MR images (b) of<br />

the male nude mouse bearing prostate carcinoma DU-145 (left) and human Kaposi’s sarcoma (SLK) tumor<br />

(right) at different time points. The arrow points to the tumor. Imaging parameters were TRZ7.8 ms,<br />

TEZ2.7 ms, 258 flip angle, and 0.4-mm coronal slice thickness.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


206<br />

11.3.2 MR IMAGING OF A PARAMAGNETIC HPMA CONJUGATE FOR BONE DELIVERY<br />

Kopecek and coworkers labeled HPMA copolymers (55 kDa) with Gd-DOTA to noninvasively<br />

evaluate macromolecular accumulation in arthritic joints in a rat model. The copolymers were<br />

labeled with fluorescein in addition to the paramagnetic metal chelate. The structure of the conjugate<br />

is shown in Figure 11.3. After intravenous administration of the conjugate in a rat model with<br />

adjuvant-induced arthritis, the biodistribution of copolymers was followed <strong>by</strong> contrast-enhanced<br />

MRI at various time points post-injection. 35<br />

Contrast-enhanced dynamic MRI clearly revealed the pharmacokinetic properties and biodistribution<br />

of the conjugate with high anatomic resolution in the adjuvant-induced arthritic rats<br />

(Figure 11.4). The copolymers had a relatively high molecular weight and a considerable<br />

amount of the copolymers was still circulating in the blood 3 h post-injection. Selective accumulation<br />

of HPMA copolymers was gradually shown in the arthritic joints of the adjuvant-induced<br />

arthritis rats. The copolymers were also cleared from the accumulation sites over time as shown <strong>by</strong><br />

contrast-enhanced MRI. The uptake and retention of the MR contrast agent labeled polymer<br />

correlated well with the histopathological features of inflammation and local tissue damages.<br />

No significant contrast enhancement was observed in the hind-limb joints of healthy rats in<br />

a control study.<br />

11.4 MR IMAGING OF PARAMAGNETIC POLY(L-GLUTAMIC ACID)<br />

CONJUGATES<br />

Poly(L-glutamic acid) (PGA) is a negatively charged, biocompatible polymer that has been used as<br />

a carrier for drug delivery. 36 It is a homopolymer of L-glutamic acid linked through peptide bonds<br />

and the pendant g-carboxyl groups are suitable to load therapeutic agents via chemical conjugation.<br />

Several poly(L-glutamic acid)-anti-cancer drug conjugates have been developed for cancer treatment.<br />

37–39 The conjugation of a MRI contrast agent to PGA enables noninvasive investigation of its<br />

pharmacokinetics and in vivo drug delivery with contrast-enhanced MRI.<br />

11.4.1 MR IMAGING OF PGA–(GD-DOTA) CONJUGATE IN AN ANIMAL TUMOR MODEL<br />

Poly(L-glutamic acid)–1,6-hexanediamine–(Gd-DOTA) conjugate was prepared using a synthetic<br />

procedure similar to that of PGA-cystamine-Gd-DOTA conjugate synthesis 40 to noninvasively<br />

HO<br />

CH 2<br />

S<br />

CH3 C<br />

m<br />

O<br />

NH<br />

CH2 CH2 CH2 NH<br />

NH<br />

O<br />

CH 2<br />

COOH<br />

O<br />

CH3 C<br />

n<br />

O<br />

NH<br />

CH2 CH OH<br />

CH3 CH 2<br />

O<br />

- OOC<br />

Nanotechnology for Cancer Therapy<br />

CH3 C<br />

O<br />

NH<br />

CH2 CH<br />

CH2 NH<br />

N<br />

Gd 3+<br />

N<br />

N N<br />

COO -<br />

COO -<br />

FIGURE 11.3 The structure of HPMA-MAFITC copolymers labeled with Gd(III)-DOTA.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Noninvasive Visualization of In Vivo Drug Delivery of Paramagnetic Polymer Conjugates 207<br />

FIGURE 11.4 The MR images of rats taken at different time points. The acquired images were post-processed<br />

using the maximum intensity projection (MIP) algorithm. P-Gd(III)-DOTA is abbreviated as P-Gd. (A-H) AIA<br />

rat images at baseline (a) and 5 min (b), 1 h (c), 2 h (d), 3 h (e), 8 h (f), 32 h (g), 43 h (h) post-injection of<br />

P-Gd(III)-DOTA. (i–n) Healthy rat images at baseline (i) and 5 min (j), 1 h (k), 2 h (l), 8 h (m), 48 h (n)<br />

post-injection of P-Gd(III)-DOTA. Arrow points to the diseased joint (Adapted from Wang, D. etal., Pharm.<br />

Res., 21, 1741, 2004. With permission.)<br />

study delivery efficiency of the conjugate. The structure of the conjugate is shown in Figure 11.5.<br />

PGA–1,6-hexanediamine–(Gd-DOTA) conjugates, H a high molecular weight (50 kDa, PDZ<br />

1.12) conjugate and a low molecular weight conjugate (20 kDa, PDZ1.08) were investigated <strong>by</strong><br />

contrast-enhanced MRI to study the size effect of conjugates on the efficiency of in vivo drug<br />

delivery. The conjugates with different molecular weights had similar Gd content (41.7% for<br />

50 kDa conjugate and 40.2 mol% for 20 kDa conjugate) and T1 relaxivity (9.44 and 9.20 mM K<br />

1 s K1 for 50 and 20 kDa conjugates, respectively). The conjugates were intravenously administered<br />

into female nu/nu athymic mice bearing human breast carcinoma MB-231 xenografts at a dose of<br />

0.07 mmol-Gd/kg. Dynamic contrast-enhanced MRI clearly revealed the size effect of the conjugates<br />

on their pharmacokinetics and in vivo tumor delivery efficiency in the animal model.<br />

Figure 11.6 shows the T1-weighted coronal MR images of mice before and at various time<br />

points after injection with the conjugates. Strong contrast enhancement was observed in the heart at<br />

the initial stage post-injection and then gradually faded away for both conjugates. The signal<br />

intensity decreased more rapidly for the low molecular weight conjugate than the high molecular<br />

weight conjugate, validating the concept that the blood circulation of the conjugates is prolonged<br />

NH<br />

CH<br />

CH2 CH2 C O<br />

OH<br />

O<br />

C<br />

m<br />

NH<br />

CH<br />

CH2 CH2 C O<br />

O<br />

C<br />

n<br />

NH(CH 2) 6NH<br />

O<br />

C<br />

− OOC<br />

N N<br />

Gd<br />

N N<br />

3+<br />

COO −<br />

COO −<br />

FIGURE 11.5 The structure of poly(L-glutamic acid)–1,6-hexanediamine–(Gd-DOTA) conjugate.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


208<br />

FIGURE 11.6 Coronal MR slice images of mice bearing MB-231 breast cancer tumors using PGA–1,6-hexanediamine–(Gd-DOTA)<br />

conjugates with different molecular weights. The images were taken at different time<br />

points using a 3D FLASH pulse sequence (1.74-ms TE, 4.3-ms TR, 258 RF tip angle, 120-mm FOV, and<br />

1.6-mm coronal slice thickness).<br />

with increase in molecular weights. The low-molecular-weight conjugate (20 kDa) was cleared<br />

from the blood circulation within 30 min after-injection. The dynamic changes in signal intensity in<br />

the liver correlated well to that in the heart. The enhancement in the liver returned to the background<br />

level 24 h after the injection, indicating low liver uptake of the PGA conjugates.<br />

The contrast enhancement pattern in the tumor tissue was similar to that of HPMA conjugates<br />

in Kaposi’s sarcoma. Strong signal was observed at tumor periphery and little in the inner tumor<br />

tissue. The dynamic MR images showed that the intensity and duration of tumor enhancement<br />

strongly depended on the size of the conjugates, indicating size-dependent tumor accumulation.<br />

Rapid clearance of the low molecular weight conjugate (20 kDa) from the blood circulation<br />

resulted in low and short accumulation in the tumor tissue. The high molecular weight conjugate<br />

(50 kDa) had a long blood circulation time, resulting in an increased contrast enhancement for at<br />

least 4 h. The enhancement also gradually extended into inner tumor tissue, indicating diffusion of<br />

the conjugate further into the tumor. The results clearly showed that drug delivery with polymeric<br />

conjugate into tumor tissue was a slow process and H conjugates with high molecular weight and<br />

long blood circulation time was more effective for tumor delivery than the conjugate with low<br />

molecular weight. Contrast-enhanced MRI clearly revealed the size effect of the conjugates on their<br />

pharmacokinetics and in vivo drug delivery efficiency.<br />

11.4.2 MR IMAGING OF PGA–MCE6–(Gd-DOTA) CONJUGATE<br />

IN AN ANIMAL TUMOR MODEL<br />

Nanotechnology for Cancer Therapy<br />

Most anti-cancer drugs are lipophilic and the conjugation of lipophilic drugs to a polymeric carrier may<br />

also modify the in vivo behavior of the polymers. Mesochlorin e 6 (Mce 6) is a lipophilic photosentizer<br />

for photodynamic therapy. 41,42 Aparamagneticpoly(L-glutamic acid) Mce6 conjugate was prepared to<br />

investigate the pharmacokinetics, biodistribution and drug delivery efficiency of the PGA–Mce6 conjugate<br />

with contrast-enhanced MRI. 43 The structure of the conjugate is shown in Figure 11.7. A PGA–1,6hexanediamine–(Gd-DOTA)<br />

conjugate was also prepared from the same poly(L-glutamic acid) as a<br />

control. Mce6 content was 2.5 mol% in PGA–Mce6–1,6-hexanediamine–(Gd-DOTA) and Gd content<br />

was 20 and 29 mol% for PGA–Mce6–1,6-hexanediamine–(Gd-DOTA) and PGA–1,6-hexanediamine–<br />

(Gd-DOTA), respectively. The T 1 relaxivity was 8.46 and 8.33 mM K1 s K1 for PGA–Mce 6–1,6-hexanediamine–(Gd-DOTA)<br />

and PGA–1,6-hexanediamine–(Gd-DOTA), respectively. The small amount<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Noninvasive Visualization of In Vivo Drug Delivery of Paramagnetic Polymer Conjugates 209<br />

NH<br />

CH<br />

CH2 CH = 2<br />

C O<br />

NH<br />

CH2 O<br />

CH2 NH<br />

O<br />

C<br />

m<br />

HOOC<br />

NH<br />

CH<br />

CH2 CH2 C O<br />

OH<br />

NH N<br />

mesoc (<br />

HOOC<br />

N<br />

HN<br />

O<br />

C<br />

n<br />

− OOC<br />

N N<br />

Gd<br />

N N<br />

3+<br />

COO −<br />

COO −<br />

of Mce6 in PGA–Mce6–1,6-hexanediamine–(Gd-DOTA) significantly reduced the hydrodynamic<br />

volume and molecular weight distribution of the conjugate. The weight (Mw) and number (Mn)<br />

average molecular weights decreased from 101 and 49 kDa of PGA–1,6-hexadiamine–(Gd-DOTA)<br />

to 49 and 34 KDa for PGA–Mce6–1,6-hexadiamine–(Gd-DOTA). The lipophilic Mce6 may alter the<br />

conformation of the polymers, resulting in a smaller hydrodynamic volume for PGA–Mce6–1,6-hexadiamine–(Gd-DOTA).<br />

A contrast-enhanced MRI study demonstrated that PGA–Mce6–1,6-hexanediamine–(Gd-DOTA)<br />

had significantly different pharmacokinetics and biodistribution from PGA–1,6-hexanediamine–(Gd-<br />

DOTA) in female nu/nu athymic mice bearing human breast carcinoma MB-231 xenografts.<br />

Figure 11.8 shows coronal dynamic MR images of mice injected with PGA–Mce6–1,6-hexanediamine–(Gd-DOTA)<br />

and PGA–1,6-hexanediamine–(Gd-DOTA) at a dose of 0.07 mmol-Gd/kg.<br />

The conjugates demonstrated different dynamic contrast enhancement patterns in the heart and liver.<br />

PGA–Mce 6–1,6-hexadiamine–(Gd-DOTA) had a strong contrast enhancement in the liver and<br />

relatively weak enhancement in the heart, whereas PGA–1,6-hexadiamine–(Gd-DOTA) exhibited<br />

NH<br />

CH<br />

CH2 CH2 C O<br />

O<br />

C<br />

NH(CH 2) 6NH<br />

FIGURE 11.7 The structure of PGA–Mce6–1,6-hexanediamine–(Gd-DOTA).<br />

FIGURE 11.8 Coronal contrast-enhanced MR slice images for PGA–Mce 6–(Gd-DOTA) conjugate and<br />

PGA–(Gd-DOTA) conjugate through the heart (b) before and at 5, 15, 30, 60, and 120 min post-injection.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

l<br />

O<br />

C


210<br />

strong enhancement in the heart and relatively weak enhancement in the liver. The conjugation of<br />

Mce 6 significantly altered the pharmacokinetics and biodistribution of the poly(L-glutamic acid). The<br />

results indicated that PGA–Mce 6–1,6-hexadiamine–(Gd-DOTA) had higher liver accumulation than<br />

PGA–1,6-hexadiamine–(Gd-DOTA). Consequently, it cleared from the blood circulation more rapidly<br />

than PGA–1,6-hexadiamine–(Gd-DOTA), which could also be attributed to the larger<br />

hydrodynamic volume of PGA–1,6-hexadiamine–(Gd-DOTA).<br />

11.5 SUMMARY<br />

Contrast-enhanced MRI is effective for noninvasive visualization of the real-time pharmacokinetics<br />

and biodistribution of paramagnetically labeled polymer drug conjugates after intravenous administration.<br />

The technique provides three-dimensional anatomic images of soft tissues with high<br />

spatial resolution. The circulation and accumulation of the paramagnetic polymeric conjugates<br />

in organs or tissues result in bright signal for T 1-weighted images. Dynamic contrast-enhanced<br />

MRI clearly reveals circulation of the conjugates in the blood, interaction with the major organs,<br />

including the liver, heart, kidneys, etc., and accumulation and clearance in the target tissues. It also<br />

provides information on the dynamic changes of the distribution patterns of polymer conjugates in<br />

solid tumor tissues, which cannot be obtained with conventional pharmacokinetic methods. Such a<br />

detailed map of the delivery system deposition within the tissue and its correlation with the local<br />

pathologic features may help to optimize the structure of the polymeric drug conjugate to achieve<br />

high drug delivery efficiency. One limitation of contrast-enhanced MRI is the accurate quantification<br />

of the concentrations of the conjugates in the tissues and organs. Several technical factors<br />

control the accurate measurement of concentrations of contrast agents with conventional contrastenhanced<br />

MRI. For quantitative study of in vivo drug delivery with paramagnetic conjugates, MR<br />

T1 mapping can be used because T1 relaxation rate (1/T1) is linearly proportional to the concentration<br />

of MRI contrast agents. Data acquisition and algorithms to process the data for T1 mapping<br />

are more complicated than for conventional contrast-enhanced MRI.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Kopecek, J. et al., HPMA copolymer-anticancer drug conjugates: design, activity and mechanism of<br />

action, Eur. J. Pharm. Biopharm., 50, 61, 2000.<br />

2. Kopecek, J. et al., Water soluble polymers in tumor targeted delivery, J. Control. Rel., 74, 147, 2001.<br />

3. Hobbs, S. K. et al., Regulation of transport pathways in tumor vessels: role of tumor type and<br />

microenvironment, Proc. Natl Acad. Sci. USA, 95, 4607, 1998.<br />

4. Maeda, H., Sawa, T., and Konno, T., Mechanism of tumor-targeted delivery of macromolecular drugs,<br />

including the EPR effect in solid tumor and clinical overview of the prototype polymeric drug<br />

SMANCS, J. Control. Rel., 74, 47, 2001.<br />

5. Minko, T., Kopeckova, P., and Kopecek, J., Chronic exposure to HPMA copolymer-bound adriamycin<br />

does not induce multidrug resistance in a human ovarian carcinoma cell line, J. Controlled Rel.,<br />

59, 133, 1999.<br />

6. Minko, T. et al., HPMA copolymer bound adriamycin overcomes MDR1 gene encoded resistance in a<br />

human ovarian carcinoma cell line, J. Control. Rel., 54, 223, 1998.<br />

7. Abe, S. and Otsuki, M., Styrene maleic acid neocarzinostatin treatment for hepatocellular carcinoma,<br />

Curr. Med. Chem. Anti-Cancer Agents, 2, 715, 2002.<br />

8. Brewerton, L. J., Fung, E., and Snyder, F. F., Polyethylene glycol-conjugated adenosine phosphorylase:<br />

development of alternative enzyme therapy for adenosine deaminase deficiency, Biochim.<br />

Biophys. Acta, 1637, 171, 2003.<br />

9. Pinheiro, J. P. et al., Drug monitoring of PEG-asparaginase treatment in childhood acute lymphoblastic<br />

leukemia and non-Hodgkin’s lymphoma, Leuk. Lymphoma, 43, 1911, 2002.<br />

10. Ettinger, L. J. et al., An open-label, multicenter study of polyethylene glycol-L-asparaginase for the<br />

treatment of acute lymphoblastic leukemia, Cancer, 75, 1176, 1995.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Noninvasive Visualization of In Vivo Drug Delivery of Paramagnetic Polymer Conjugates 211<br />

11. Motzer, R. J. et al., Phase II trial of branched peginterferon-alpha 2a (40 kDa) for patients with<br />

advanced renal cell carcinoma, Ann. Oncol., 13, 1799, 2002.<br />

12. Rowinsky, E. K. et al., A phase I and pharmacokinetic study of pegylated camptothecin as a 1-hour<br />

infusion every 3 weeks in patients with advanced solid malignancies, J. Clin. Oncol., 21, 148, 2003.<br />

13. Auzenne, E. et al., Superior therapeutic profile of poly-L-glutamic acid-paclitaxel copolymer<br />

compared with Taxol in xenogeneic compartmental models of human ovarian carcinoma, Clin.<br />

Cancer Res., 8, 573, 2002.<br />

14. Gianasi, E. et al., HPMA copolymer platinates as novel antitumor agents: in vitro properties, pharmacokinetics<br />

and antitumor activity in vivo, Eur. J. Cancer, 35, 994, 1999.<br />

15. Thomson, A. H. et al., Population pharmacokinetics in phase I drug development: a phase I study of<br />

PK1 in patients with solid tumors, Br. J. Cancer, 81, 99, 1999.<br />

16. Seymour, L. W. et al., Hepatic drug targeting: phase I evaluation of polymer-bound doxorubicin,<br />

J. Clin. Oncol., 20, 1668, 2002.<br />

17. Liang, Z. P. and Lauterbur, P. C., Principles of Magnetic Resonance Imaging, IEEE Press, New York,<br />

1999.<br />

18. Vlaardingerbroek, M. T. and den Boer, J. A., Magnetic Resonance Imaging, 3rd ed., Springer, New<br />

York, 2003.<br />

19. Lauffer, R. B., Paramagnetic metal complexes as water proton relaxation agents for NMR imaging:<br />

theory and design, Chem. Rev., 87, 901, 1987.<br />

20. Merbach, A. E. and Toth, E., The Chemistry of Contrast Agents in Medical Magnetic Resonance<br />

Imaging, John Wiley & Sons, Inc., England, 2001 chap. 1–2<br />

21. Weinmann, H. J. et al., Characteristics of gadolinium-DTPA complex: a potential NMR contrast<br />

agent, Am. J. Roentgenol., 142, 619, 1984.<br />

22. Kaplan, G. D., Aisen, A. M., Aravapalli, S. R. et al., Preliminary clinical trial of gadodiamide<br />

injection: a new nonionic gadolinium contrast agent for MR imaging, J. Magn. Reson. Imaging, 1,<br />

57, 1991.<br />

23. Magerstadt, M. et al., Gd(DOTA): An alternative to Gd(DTPA) as a T1,2 relaxation agent for NMR<br />

imaging or spectroscopy, Magn. Reson. Med., 3, 808, 1986.<br />

24. Tweedle, M. F., The proHance story: the making of a novel MRI contrast agent, Eur. Radiol.,<br />

7(Suppl 5), 225, 1997.<br />

25. Youk, J. H., Lee, J. M., and Kim, C. S., MRI for detection of hepatocellular carcinoma: comparison of<br />

mangafodipir trisodium and gadopentetate dimeglumine contrast agents, Am. J. Roentgenol., 183,<br />

1049, 2004.<br />

26. Weissleder, R. et al., Ultrasmall superparamagnetic iron oxide: an intravenous contrast agent for<br />

assessing lymph nodes with MR imaging, Radiology, 175, 494, 1990.<br />

27. Wang, S. C. et al., Evaluation of Gd-DTPA-labeled Dextran as an intravascular MR contrast agent:<br />

imaging characteristics in normal rat tissues, Radiology, 175, 483, 1990.<br />

28. Bryant, L. H. et al., Synthesis and relaxometry of high-generation (GZ5, 7, 9 and 10) PAMAM<br />

dendrimer-DOTA-gadolinium chelates, J. Magn. Reson. Imaging, 9, 348, 1999.<br />

29. Lu, Z. R. et al., Design of novel bioconjugates for targeted drug delivery, J. Control. Rel., 78, 165,<br />

2002.<br />

30. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2, 347, 2003.<br />

31. Pimm, M. V. et al., Gamma scintigraphy of the biodistribution of 123 I labeled N-(2-hydroxypropyl)methacrylamide<br />

copolymer-doxorubicin conjugates in mice with transplanted melanoma and<br />

mammary carcinoma, J. Drug Target, 3, 375, 1996.<br />

32. Kissel, M. et al., Synthetic macromolecular drug carriers: biodistribution of poly [(N-2-hydroxypropyl)methacrylamide]<br />

copolymers and their accumulation in solid rat tumors, J. Pharm. Sci. Tech., 55,<br />

191, 2001.<br />

33. Mitra, A. et al., Technetium-99m-Labeled N-(2-hydroxypropyl) methacrylamide copolymers:<br />

synthesis, characterization, and in vivo biodistribution, Pharm. Res., 21, 1153, 2004.<br />

34. Wang, Y.L., and Lu, Z.R. Unpublished data 2005.<br />

35. Wang, D. et al., The arhrotropism of macromolecules in adjuvant-induced arthritis rat model: a<br />

preliminary study, Pharm. Res., 21, 1741, 2004.<br />

36. Li, C., Poly (L-glutamic acid)-anticancer drug conjugates, Adv. Drug Deliv. Rev., 54, 695, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


212<br />

Nanotechnology for Cancer Therapy<br />

37. Hurwitz, E., Wilchek, M., and Pitha, J., Soluble macromolecules as carriers for daunorubicin, J. Appl.<br />

Biochem., 2, 25, 1980.<br />

38. Li, C. et al., Complete regression of well-established tumors using novel water-soluble poly (Lglutamic<br />

acid)-paclitaxel conjugates, Cancer Res., 58, 2404, 1998.<br />

39. Zou, Y. et al., Effectiveness of water soluble poly(L-glutamic acid)–camptothecin conjugate against<br />

resistant human lung cancer xenografted in nude mice, Int. J. Oncol., 18, 331, 2001.<br />

40. Lu, Z. R. et al., Poly(L-glutamic acid) Gd(III)-DOTA conjugate with a degradable spacer for magnetic<br />

resonance imaging, Bioconjugate Chem., 14, 715, 2003.<br />

41. Lu, Z. R., Kopeckova, P., and Kopecek, J., Polymerizable Fab’ fragment for targeting of anticancer<br />

drugs, Nature Biotechnol., 17, 1101, 1999.<br />

42. Lu, Z. R. et al., Polymerizable Fab’ antibody fragments targeted photodynamic cancer therapy in nude<br />

mice, STP Pharm. Sci., 13, 69, 2003.<br />

43. Vaidya, A. et al. Non-invasive in vivo imaging of a polymeric drug conjugate using contrast enhanced<br />

MRI. In Transactions of 32nd Annual Meeting of the Controlled Release Society, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


12<br />

CONTENTS<br />

Polymeric Nanoparticles for<br />

Tumor-Targeted Drug Delivery<br />

Tania Betancourt, Amber Doiron,<br />

and Lisa Brannon-Peppas<br />

12.1 Introduction ........................................................................................................................215<br />

12.2 Targeting to Cancer............................................................................................................216<br />

12.3 Passive Targeting and the EPR Effect ...............................................................................217<br />

12.4 Targeting to Angiogenesis .................................................................................................218<br />

12.4.1 Targeting Using Vascular Endothelial Growth Factor Receptors...................... 218<br />

12.4.2 Targeting Using Integrins ................................................................................... 218<br />

12.4.2.1 Integrins as Targets for Imaging .........................................................220<br />

12.5 Targeting Using Folate Receptors .....................................................................................220<br />

12.5.1 Antibodies and Folate Receptors ........................................................................ 221<br />

12.5.2 Folate-Targeted Nanoparticles for Gene Delivery ............................................. 222<br />

12.6 Approaches for Cancer Targeting to Specific Cancer Types ............................................223<br />

12.6.1 Prostate Cancer.................................................................................................... 224<br />

12.7 Targeted Nanoparticles and Imaging of Cancer................................................................225<br />

12.8 Other Targets for Cancer ...................................................................................................225<br />

12.9 Avidin and Biotin Targeting ..............................................................................................226<br />

12.10 Conclusions ........................................................................................................................226<br />

References.....................................................................................................................................226<br />

12.1 INTRODUCTION<br />

Cancer is a disease that affects millions of people across the globe every year. The World Health<br />

Organization estimated that more than 10 million people developed a malignant tumor and more<br />

than 6.5 million people died from this disease during the year 2000. 1 In the United States, cancer is<br />

the second cause of deaths from disease after heart disease, accounting for more than half a million<br />

deaths every year. According to the American Cancer Society (ACS) cancer statistics, the overall<br />

cost for cancer for the United States in 2004 was $189.8 billion: $69.4 billion for direct medical<br />

costs, $16.9 billion for indirect morbidity costs, and $103.5 billion for indirect mortality costs. 2<br />

Furthermore, while mortality rates of other major chronic diseases, such as heart and cerebrovascular<br />

disease, decreased significantly in the past half-century, cancer mortality rates have<br />

remained approximately constant. 2 This is a troubling fact because it suggests that recent detection<br />

and treatment options have not been able to improve mortality rates substantially.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

215


216<br />

Research in the past decade has focused on using unique characteristics of cancer cells and the<br />

vasculature surrounding those cells to deliver imaging agents, chemotherapeutic drugs, gene<br />

therapy, and other active agents directly and selectively to cancerous tissues. Many of these new<br />

formulations (described elsewhere in this volume) are liposomes, prodrugs, polymer conjugates,<br />

micelles, and dendritic systems. This chapter will concentrate on polymeric nanoparticles that have<br />

been studied as targeted systems for treatment and detection of cancer.<br />

12.2 TARGETING TO CANCER<br />

Nanotechnology for Cancer Therapy<br />

Nanoparticles may be targeted to the growing vasculature serving the growing cancer or to the<br />

cancer cells themselves. Targeted delivery utilizes unique phenotypic features of diseased tissues<br />

and cells in order to concentrate the drug at the location where it is needed. Targeted delivery can be<br />

divided into passive and active targeting. Passive targeting tries to minimize nonspecific<br />

interactions between the drug carrier and nontarget sites in the body <strong>by</strong> detailing the physiochemical<br />

properties of the aberrant tissue such as size, morphology, hydrophilicity, and surface charge. 3<br />

When targeting tumor tissue, the enhanced permeability and retention effect (EPR) is an example of<br />

passive targeting approach; it allows passage of drug carriers ranging in size from 10 to 500 nm<br />

through the highly permeable blood vessels that supply growing tumors and leads to entrapment of<br />

large molecules as a result of deficient lymphatic drainage. 3–5 In fact, it has been reported that the<br />

intra-cellular openings in vascular endothelium of tumor blood vessels can be up to 2 mm in<br />

diameter and that the vessel leakiness in tumor vasculature can be up to an order of magnitude<br />

higher than that of normal blood vessels. 5 Active targeting utilizes biologically specific interactions<br />

including antigen–antibody and ligand–receptor binding and may seek drug uptake <strong>by</strong> receptormediated<br />

endocytosis through association of the drug or drug carrier with such antigen or ligand. 3<br />

Receptor-mediated endocytosis commonly occurs through clathrin-coated vesicles and is carried<br />

out in mammalian cells continuously for the uptake of nutrients and for modulation of signal<br />

transduction through the up- or down-regulation of signaling receptors. 6 Targeted delivery<br />

avoids the need for high systemic drug levels for the drug to be effective and consequently<br />

offers a more economic alternative for treatment. Not only is it useful for therapeutic purposes;<br />

it is also beneficial in diagnosis. Recent research has pointed to its ability to concentrate imaging or<br />

contrast agents for the detection of malignancies and for monitoring the effects of therapeutic<br />

agents. 7,8 To date, most systems for targeted delivery have utilized drug conjugates, liposomes<br />

or micelles. 9–11 Targeting of particulate systems has focused more often on passive targeting based<br />

on size than on active targeting. But systems that combine both methods, starting with passive<br />

targeting through EPR and enhancing the targeting through specific interactions are beginning to<br />

show great promise.<br />

While it is challenging to deliver a drug or imaging agent-containing nanoparticle directly and<br />

selectively to a cancerous cell or tissue, the additional challenges of then having that particle and/or<br />

its contents being transported into the targeted cell have often been overlooked. Couvreur presented<br />

some of these challenges at the 11th International Symposium on Recent Advances in Drug<br />

Delivery Systems and also summarized that presentation in a recent publication. 12 The tumor<br />

resistance can be due to the deliverance of nanoparticles to the tumor as well as to resistance to<br />

the active agent being delivered. In this article, many different pathways are described and the<br />

enhancement of drug delivery due to the presence of nanoparticles, especially polycyanoacrylate<br />

nanoparticles, is evaluated and summarized. The enhancement of drug permeability into cells due<br />

to interactions with biodegradation <strong>by</strong>products as well as the effect of nanoparticle surface charge<br />

is discussed.<br />

In this chapter, we will first review recent work on targeting to the two most promising targets<br />

for cancer: Angiogenesis and folate receptors. We will then describe other potential targets for<br />

cancer imaging and therapy with nanoparticles, including antibodies strategies using biotin.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 217<br />

12.3 PASSIVE TARGETING AND THE EPR EFFECT<br />

Although active targeting may be achieved through targeting to folate receptors, angiogenesis, or<br />

targets more specific to the cancer being treated, some enhancement of treatment to cancers on<br />

account of the EPR effect cannot be ignored and should be utilized until more specific targeting<br />

systems can be developed. Preparation of nanoparticle systems that can avoid uptake <strong>by</strong> the<br />

reticuloendothelial system (RES) is essential, and such particles are often referred to as<br />

“stealth” nanoparticles.<br />

The most effective polymer used as a coating on nanoparticles to avoid detection <strong>by</strong> the RES is<br />

poly(ethylene glycol). 13 The latter can be achieved <strong>by</strong> adsorption of the PEG-containing polymer<br />

onto the surface of nanoparticles, direct conjugation of the PEG to the nanoparticles, or inclusion of<br />

PEG in the polymeric backbone that makes up the nanoparticles. The PEG itself may then also be<br />

modified to give targeting capabilities and to avoid uptake <strong>by</strong> the RES. The PEG works to mask the<br />

surface of the nanoparticles <strong>by</strong> reducing the plasma and protein adsorption to the particles, reducing<br />

the complement activation and hence the recognition of the PEG as a foreign substance in the blood<br />

stream. The longer circulation time afforded PEGylated nanoparticles allows them time to be<br />

targeted, whether passively or actively, to cancerous tissues.<br />

Extended circulation time and enhanced tumor targeting were seen for poly(ethylene oxide)modified<br />

poly(epsilon-caprolactone) nanoparticles in mice with tumors of MDA-MB-231, a human<br />

breast carcinoma. 14 These particles were loaded with [ 3 H]-tamoxifen. The amount of labeled<br />

tamoxifen at the tumor site, 6 h after injection, for those particles with PEO modification was at<br />

least twice that of particles without modification and four times that of labeled tamoxifen injection.<br />

The amount of labeled tamoxifen found in the blood stream at 6 h after injection was also at least<br />

twice that of the injection or unmodified nanoparticle formulations.<br />

The stability and circulation of PLGA–mPEG nanoparticles containing cisplatin was investigated.<br />

It was found that while the mPEG content affected the drug release rate, the drug loading<br />

level had no effect on the drug release rate for in vitro studies. 15 The release was more that 60%<br />

completed within the first 12 h in all cases. Data for blood levels was only presented for 3 h, so it is<br />

hard to draw any valid conclusions from this information.<br />

Nanoparticles of PLA and PEG–PPG–PEG were prepared containing irinotecan, a prodrug of<br />

an analogue of camptothecin. 16 Although little characterization beyond the average particles size<br />

(231 nm) was presented here and no in vitro studies were described, the in vivo studies are quite<br />

interesting. In this study, there was a modest increase in survival time in mice with M5076 tumors<br />

(early liver metastatic stage) after a single injection and more pronounced increases in survival time<br />

after either two or three repeat injections. It is noteworthy that the greatest survival times (20%<br />

survival at 45 days at the end of the study) were seen with two injections at days three and five after<br />

the implantation of the tumor.<br />

Polycyanoacrylate nanoparticles have long been studied <strong>by</strong> Couvreur and collaborators as<br />

biodegradable nanoparticles for a variety of applications and therapies which are not limited to<br />

treatment of cancer. 17 A recent work describes the effectiveness of these nanoparticles as delivery<br />

systems for brain tumor targeting. Here they studied uncoated and PEG-coated nanoparticles and<br />

found that both types of particles showed accumulation in a well-established 9L gliosarcoma in rat<br />

studies. The PEG-coated particles showed the highest accumulation, with a tumor-to-brain ratio<br />

of 11.<br />

Poly(butyl cyanoacrylate) nanoparticles containing doxorubicin were prepared with no surface<br />

modifications; the in vivo distribution of 99m Tc labeled nanoparticles was evaluated in mice inoculated<br />

with Dalton’s lymphoma tumor cells. 18 The nanoparticles were administered <strong>by</strong> subcutaneous<br />

injection and the concentration in a number of organs was followed for 48 hours and compared with<br />

that for 99m Tc labeled doxorubicin alone. When compared with the amount of doxorubicin alone,<br />

the amount of radioactivity in the tumor was higher at all times tested, with a 13-fold increase seen<br />

at 48 h.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


218<br />

While the majority of work on targeted nanoparticles has been carried out with polylactides,<br />

polyglycolides and polycyanoacrylates, those are not the only materials that can be used in nanoparticles.<br />

Some recent work with radiolabeled gelatin nanoparticles modified with poly(ethylene<br />

glycol) showed that adding the PEG to the surface of these nanoparticles increased the circulation<br />

time in mice with double the amount of PEG-modified nanoparticles in the blood stream 3 h after<br />

injection compared to the amount of control nanoparticles. 19 In addition, there was a four-fold<br />

increase in the amount of PEG-modified gelatin nanoparticles found in the tumor 4 h after injection<br />

and later when compared to the number of nonmodified gelatin nanoparticles.<br />

12.4 TARGETING TO ANGIOGENESIS<br />

A recently explored and potentially promising target of cancer drug and gene nanoparticle therapy<br />

is tumor angiogenesis. It is now well established that tumor growth is dependent on new capillary<br />

infiltration from surrounding, preexisting vasculature. 20,21 This is an important control point in<br />

cancer as much research has proven that tumors cannot effectively grow past a small size or<br />

metastasize without blood supply. 22–24 Except for the cases of menstruation, wound healing, and<br />

tissue regeneration, capillaries do not increase in size or number under normal physiological<br />

conditions. Tumor growth is an exception to this physiological rule.<br />

Tumors are typically unable to affect angiogenesis when they are small and surrounded <strong>by</strong><br />

healthy tissue. However, at the point in growth where nutrients, oxygen, and growth factors can no<br />

longer reach the cancer cells, blood flow is required to allow further growth of the tumor. After what<br />

is occasionally a substantial time period, the tumor may abruptly induce angiogenesis into the<br />

tissue. 23 Because understanding this step in the progression of cancer is thought to be of great<br />

importance, much research has been and continues to be focused on pinpointing the progression of<br />

cancer and on targeting the event for therapy.<br />

Much research has focused on targeted therapy of either chemotherapeutic agents to sites of<br />

tumor angiogenesis or of angiogenesis-inhibiting drugs to tumors with the goal of directly combatting<br />

the proliferation of newly forming capillaries in the tumor. Angiogenesis is a complex, multicomponent<br />

process that involves many cell types, cytokines, growth factors and receptors,<br />

proteases, and adhesion molecules. 25 As a result, there are many potential targets for anti-angiogenic<br />

or chemotherapeutic therapy. Some recent advances in targeting approaches for nanoparticle drug<br />

delivery are discussed below.<br />

12.4.1 TARGETING USING VASCULAR ENDOTHELIAL GROWTH FACTOR RECEPTORS<br />

Vascular endothelial growth factor (VEGF) is particularly important in the process of angiogenesis<br />

and has been shown to greatly affect tumor growth in animal models. 26–28 The VEGF receptors have<br />

been used as a means to target the vascular bed in many instances. VEGF receptor-2 (VEGFR-2)<br />

has recently been used to target nanoparticles to tumor vascular beds <strong>by</strong> Li et al in mice with<br />

K1735-M2 tumors. 29 A succinyl-dextran-polymerized nanoparticle conjugated to rat anti-mouse<br />

VEGFR-2 antibody and radioisotope 90 Y caused a significant tumor growth delay compared to<br />

conventional radiolabeled antibody and other controls. Additionally, anti-CD31 staining showed<br />

a decreased vessel density and damage to tumor vessels after treatment with the anti-VEGFR-<br />

2- 90 Y nanoparticles.<br />

12.4.2 TARGETING USING INTEGRINS<br />

Nanotechnology for Cancer Therapy<br />

The integrins represent another important cell surface molecule group for angiogenesis targeting<br />

because some integrins, such as avb3 and avb5, are upregulated on the endothelial cell surface of<br />

neovasculature. 30,31 The a vb 3 integrin is expressed on numerous tumor cell types; it is highly<br />

expressed on neovascular endothelial cells. Hood and collaborators used 40-nm diameter<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 219<br />

cationic-lipid-based nanoparticles coupled to an organic avb3 ligand that was shown to be specific<br />

for avb3 in cell studies. These nanoparticles, which contained a luciferase reporter plasmid, were<br />

injected into mice with the avb3-negative cell line M21-L. The nanoparticles targeted to the<br />

neovasculature within the tumor (but not to the tumor cells themselves) with no expression elsewhere<br />

in the mice as detected <strong>by</strong> luciferase expression. In order to test therapeutic efficacy, NPs<br />

were conjugated to a mutant Raf gene that blocks angiogenesis. Systemic injection in the M21-L<br />

tumor-expressing mice rapidly induced apoptosis of endothelial cells within the tumor. Tumor<br />

regression was seen within 10 days. 32<br />

In the same study with VEGFR-2 targeting discussed above, Li and collaborators also demonstrated<br />

targeting of the radioisotope 90 Y using nanoparticles targeted to the integrin avb3 with a<br />

small molecule integrin agonist. 29 In mice with K1735-M2 tumors, avb3-targeted 90 Y-nanoparticles<br />

significantly delayed tumor growth compared to untreated tumors. TUNEL staining of tumor<br />

sections showed widespread apoptosis in tumors treated with these targeted nanoparticles. The<br />

authors have postulated that this targeted nanoparticle radiotherapy has the potential to be used<br />

to treat a variety of solid tumors. They have also postulated that the use of nanoparticles increases<br />

efficacy due to the high payload delivered <strong>by</strong> the carriers.<br />

Additionally, PEGylated polyethyleneimine (PEI) nanoplexes with a cyclic disulfide bond<br />

constrained Arg–Gly–Asp (RGD) peptide ligand at the distal end have been used to target<br />

integrin-expressing tumor neovasculature to deliver siRNA. Integrins are receptors for extracellular<br />

matrix components that contain a tripeptide RGD sequence. Therefore, RGD containing peptide<br />

sequences can target to cell surface integrins that are upregulated on neovasculature. The siRNA<br />

used inhibited angiogenesis <strong>by</strong> inhibiting VEGFR-2 expression. Intravenous injection of nanoparticles<br />

into nude mice with N2A tumors showed tumor uptake of the siRNA, inhibition of protein<br />

synthesis in the tumor, and inhibition of angiogenesis and tumor growth. 33 This study demonstrates<br />

tumor selective delivery through both the targeting ligand and gene pathway <strong>by</strong> using siRNA.<br />

In another a vb 3 targeting approach using an RGD peptide, Kopelman has created multifunctional<br />

nanoparticles of 30–60 nm for the treatment of gliomas. 34 The nanoparticles are able to kill<br />

cancer cells <strong>by</strong> bombarding them with externally released reactive oxygen species created <strong>by</strong><br />

photodynamic agents activated <strong>by</strong> laser light. The particles also contain superparamagnetic iron<br />

oxide and enhance imaging <strong>by</strong> magnetic resonance. The photodynamic sensitizer and MRI contrast<br />

agents are entrapped within a polyacrylamide core, the surface of which is coated with PEG chains<br />

and targeting RGD moieties. The particles containing photodynamic agents were shown to produce<br />

sufficient singlet oxygen to kill cells in vitro. Additionally, these nanoplatforms were injected into<br />

an in vivo rat intracerebral 9L tumor model, and diffusion MRI was performed at various times to<br />

evaluate the tumor diffusion, tumor growth, and tumor load. The gliomas treated with the nanoparticles<br />

and irradiated with laser light caused regional necrosis and significant shrinkage of tumor<br />

mass, a shrinkage that lasted for 12 days. The authors postulate that the light activated release of<br />

reactive oxygen from photosensitizer-containing nanoparticles is a viable approach for brain tumor<br />

treatment. Also, the incorporation of MRI contrast agents allows for monitoring of treatment and<br />

tumor progression in vivo.<br />

Carbohydrate based nanoparticles have also been used to target drugs to neovasculature via an<br />

avb3/cyclic RGD peptide interaction. Inulin multi-methacrylate formed the core of the nanoparticles<br />

and was attached to the RGD targeting moiety with a PEG linker. Doxorubicin was loaded in<br />

the nanoparticles via covalent and noncovalent linkages. The pharmacokinetics and biodistribution<br />

of the doxorubicin loaded nanoparticles were studied over five days in female Balb/cJ mice with<br />

metastatic mammary tumor clone-66. A bi-exponential fix with a terminal half-life of 5.99 h was<br />

observed; decreasing drug concentrations with time in the heart, lungs, kidney, and plasma was also<br />

observed. Conversely, increasing drug accumulation was observed in the liver, spleen, and in the<br />

tumor where there was also the presence of high levels of doxorubicin metabolite. The presence of<br />

the high metabolite levels in the tumorsuggests nothing more than tumor-specific nanoparticle<br />

degradation and release of drug. 35<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


220<br />

12.4.2.1 Integrins as Targets for Imaging<br />

In an effort to image angiogenesis, Mulder and collaborators have created MR-detectable and<br />

fluorescent liposomes that are targeted to avb3 integrin on neovasculature. 36 The MR contrast<br />

agent Gd-DTPA-bis(stearylamide) was incorporated into PEGylated liposomes covalently<br />

coupled to RGD peptide. It was observed that the liposomes effectively brought about an increase<br />

in signal on T1-weighted images. Injection of these lipidic nanoparticles in nude mice with subcutaneously<br />

implanted LS174 human colon adenocarcinoma led to the ability to specifically image the<br />

vascular endothelial cells in the tumor. Ex vivo fluorescence confirmed that RGD liposomes<br />

specifically interacted with tumor endothelium associated with neovasculature.<br />

In a study reported <strong>by</strong> Lanza, paramagnetic molecular imaging of angiogenesis was accomplished<br />

in vivo with avb3-targeted lipid encapsulated perfluorocarbon nanoparticles of about<br />

250 nm. 37 A Vx-2 carcinoma tumor in New Zealand rabbits was imaged with a 1.5-T MRI<br />

system with either nontargeted or avb3-targeted nanoparticles. An eight-fold greater contrast<br />

enhancement was achieved with the targeted nanoparticles. In the second part of the study, paclitaxel<br />

loaded nanoparticles were targeted to tissue factor (TF) proteins on vascular smooth muscle<br />

cells with a specific TF antibody. The TF-targeted paclitaxel particles inhibited cell proliferation<br />

while delivery of nanoparticles to targeted cells was confirmed with fluorine spectroscopy.<br />

12.5 TARGETING USING FOLATE RECEPTORS<br />

Nanotechnology for Cancer Therapy<br />

During the past few decades, there has been great interest in the utilization of folate receptors for the<br />

targeted delivery of therapeutic and imaging agents. A number of delivery systems have been<br />

utilized for this purpose, including drug conjugates, 38–40 liposomes, 41 micelles, 42 viral vectors, 43<br />

and nanoparticles.<br />

Folate (vitamin B-9) is essential for the synthesis of nucleotides and amino acids. Two main<br />

groups of molecules are responsible for transport of folate molecules in vivo. Most cells in the body<br />

express a folate anion transporter with micro-molar affinities for folates that participates in the<br />

transport of coenzyme 5-methyltetrahydrofolate, the physiologic circulating reduced form of folate.<br />

Folate receptors (FR), <strong>by</strong> contrast, are members of the glycosylphosphatidylinositol (GPI)-linked<br />

membrane glycoprotein family and have high affinity for folic acid, an oxidized form of folate, and<br />

5-methyltetrahydrofolate, with binding affinities being in the nanomolar range (K D !1!10 K9 M)<br />

for the a isoform of FR. 44–46 It has been observed with few exceptions that only cells involved in<br />

pathologic conditions, including cancer cells, express the high affinity folate receptors. These<br />

receptors are able to transport folic acid, folate-bound molecules, and even particles through<br />

receptor-mediated endocytosis. 47,48<br />

FR are known to be overexpressed in various epithelial cancer cells, such as those of ovarian,<br />

mammary gland, colon, lung, prostate, and brain epithelial cancers, and in leukemic cells. 49–61<br />

Folate receptor overexpression has been correlated to poor prognosis. In addition, metastasized<br />

cancer cells have been found to overexpress the folate receptor to a larger degree than localized<br />

tumor cells. 62 This finding is of great importance. The only nonpathological tissues where FR is<br />

expressed are choroid plexus, placenta, lungs, thyroid, and kidney. 46,63 FR expression is limited to<br />

the apical (luminal) side of polarized epithelial cells, except for the cells of the proximal tubules in<br />

the kidney. As a consequence, FR is practically inaccessible to blood-borne folate-linked<br />

systems. 44,45 These characteristics make folate receptors very advantageous for targeted delivery<br />

of nanoparticles with high payloads of therapeutic agents, imaging agents, and even genes for<br />

the treatment, detection, and monitoring of cancer. What is more, the macromolecular size of<br />

nanoparticles will prevent gromerular filtration and the consequent exposure of kidney tissue to<br />

folate-targeted nanoparticles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 221<br />

12.5.1 ANTIBODIES AND FOLATE RECEPTORS<br />

The most common targeting moieties utilized for FR targeting include monoclonal antibodies to FR<br />

and folic acid. 46 To date, monoclonal antibodies to FR have not been utilized for targeted delivery<br />

of nanoparticles although successful experiences have been reported for radiopharmaceuticals. 64<br />

The benefits of targeting with folic acid are many: it is small in size, has high stability, lacks<br />

immunogenicity, and costs very little. 62 Folic acid conjugation at its g-carboxyl group is necessary<br />

for maintaining binding affinity to FR. The structure of folic acid is shown in Figure 12.1. Cellular<br />

uptake of drug carriers bound to folic acid is believed to be mediated <strong>by</strong> receptor mediated<br />

endocytosis although this process is not currently completely understood. A number of hypotheses<br />

have been proposed for this process, including clathrin and caveolar pathways. 63 It has been shown<br />

that folate-bound molecules are able to escape endosomes after receptor-mediated endocytosis<br />

because the process of endosomal acidification results in a conformational change in the receptor<br />

that facilitates folate ligand release. Consequently, folate-bound molecules provides a great opportunity<br />

for the delivery of pH-sensitive biopharmaceuticals. 45,62<br />

In most drug delivery systems investigated thus far, folic acid is incorporated to the drug<br />

delivery system through conjugation to a poly(ethylene glycol) (PEG) spacer utilizing wellknown<br />

dicyclohexylcarbodiimide/N-hydroxysuccinimide (DCC/NHS) mediated chemistry. Such<br />

design aims to minimize steric hindrance for optimal folate recognition. This conjugation technique,<br />

however, can activate both the g- and the a-carboxylic acids of folic acid. It should be said,<br />

though, that the g group is more reactive and is responsible for most of the linkages.<br />

Numerous attempts at nanoparticle targeting to the folate receptor have been reported. Many<br />

groups have formulated nanospheres of amphiphilic block copolymers including PEG. Park and<br />

collaborators reported in vitro results of the preparation and evaluation of methoxy poly(ethylene<br />

glycol) (PEG)-poly(3-caprolactone) (PCL) block copolymer nanospheres loaded with paclitaxel in<br />

which folic acid was conjugated to a modified amino-terminated PCL with a carbodiimidemediated<br />

reaction. 65 Dialysis was used to create these nanospheres that ranged in diameter from<br />

50 to 120 nm, depending on the ratio of the block copolymers. Paclitaxel loading efficiencies of up<br />

to 55% were reported with this system, thus significantly increasing the effective solubility of this<br />

agent in aqueous systems like the body. Because the folate moiety was conjugated to the hydrophobic<br />

end of the block copolymer, it is expected that upon nanosphere formation it will<br />

be localized in the inner core of the particles. However, XPS characterization demonstrated the<br />

presence of nitrogen-containing molecules at the surface which could only be attributed to the<br />

folate linker. Although the targeting effectiveness of these particles was not determined, cytotoxicity<br />

studies revealed that encapsulation of paclitaxel into the nanospheres reduced its<br />

O<br />

OH<br />

HO<br />

O<br />

O<br />

NH<br />

FIGURE 12.1 Chemical structure of folic acid. Conjugation to folic acid for folate targeting is commonly<br />

done at the g-carboxyl group.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

NH<br />

N<br />

HO<br />

N<br />

N<br />

N<br />

NH 2


222<br />

cytotoxicity. Consequently, encapsulation offered a safe alternative to direct administration of this<br />

chemotherapeutic agent. Further studies should evaluate whether conjugation of folic acid to<br />

hydrophobic end of the block copolymer results in sufficient surface availability of this<br />

targeting agent.<br />

Another approach to folate targeting of nanoparticles has been the modification of surface<br />

properties of polymeric nanoparticles with polymer conjugates including the targeting agent. For<br />

example, Kim reported on the modification of anionic poly(lactic-co-glycolic acid) (PLGA) nanoparticles<br />

prepared <strong>by</strong> emulsification and solvent evaporation with a cationic poly(L-lysine)poly(ethylene<br />

glycol)-folate (PLL–PEG–FOL) copolymer. 66 In this design, the polycation PLL<br />

block attaches to the PLGA nanoparticle through ionic interactions. Surface coating is achieved<br />

<strong>by</strong> simple incubation in an aqueous solution containing the PLL–PEG–FOL copolymer. The PEGfolate<br />

end of the copolymer is oriented toward the outer aqueous phase for better interaction<br />

between folate and the targeted cell membrane receptor. XPS characterization demonstrated the<br />

presence of nitrogen from folic acid on the surface of the coated nanoparticles. In vitro cellular<br />

studies with FITC-labeled nanoparticles revealed an increase in uptake of coated nanoparticles with<br />

increased conjugate-to-nanoparticle ratio in KB cells. Since a decrease in uptake was seen upon<br />

addition of free folic acid in the medium, the transport of nanoparticles into the cells was attributed<br />

to endocytosis mediated <strong>by</strong> the folate receptor.<br />

Dendritic polymer systems of polyamidoamine (PAMAM) with folic acid as the targeting agents<br />

and drug and imaging agent (methotrexate or tritium and fluorescein or 6-carboxytetramethylrhodamine)<br />

were prepared and tested in mouse models. 67 It was found that targeted systems, as opposed<br />

to nontargeted systems, slowed the rate of tumor growth and even showed a complete cure in one<br />

mouse. This study was conducted with twice-weekly tail vein injections. The biodistribution studies<br />

showed that, for a single targeted injection of nanoparticles, a very high amount of the nanoparticles<br />

accumulated within the tumor <strong>by</strong> day 1; this level remained high through at least day 4.<br />

Nanometric particles prepared from drug-polymer conjugates have also been reported. In one<br />

notable study, Yoo and collaborators reported on the formulation of doxorubicin-PEG-folate nanoaggregates<br />

encapsulating additional doxorubicin in their core. 11 These aggregates are formed<br />

spontaneously when an organic phase containing the copolymer and solubilized doxorubicin is<br />

dispersed into an aqueous phase containing triethylamine. The basic aqueous environment results in<br />

deprotonation of doxorubicin, causing it to form aggregates with the hydrophobic copolymer. The<br />

average aggregate diameter was approximately 200 nm. In vitro cellular studies showed increased<br />

uptake of the nano aggregates in cells expressing the folate receptor when folic acid was absent<br />

from cell media and increased cytotoxicity (anti-tumor efficacy) of the aggregates compared to the<br />

free drug in cells expressing the FR. In vivo studies in mouse xenografts (KB cells) showed that<br />

the nanoaggregates had superior therapeutic efficacy than both doxorubicin aggregates without the<br />

folic acid ligand and free doxorubicin in solution.<br />

Nanoparticles of temperature-responsive hydrogels conjugated to folic acid have also<br />

been studied with the purpose of delivering chemotherapeutic agents. Nayak reported on poly<br />

(N-isopropylacrylamide) (pNIPAM) nanoparticles that exhibit lower critical solution temperature<br />

(LCST) behavior. 68 Fluorescent agents were included in the core of these nanoparticles to facilitate<br />

with tracking; amine comonomers were incorporated into the other pNIPAM shell for conjugation<br />

to folic acid. These nanoparticles swell when their temperature falls under their LCST. Indeed, the<br />

nanoparticle size increased from w50 nm at 378Ctow135 nm at 258C. In vitro cell uptake studies<br />

showed a 10-fold increase in the intake of nanoparticles conjugated to folic acid compared to those<br />

without the targeting agent in KB cells (FRC).<br />

12.5.2 FOLATE-TARGETED NANOPARTICLES FOR GENE DELIVERY<br />

Nanotechnology for Cancer Therapy<br />

The use of folate-targeted nanoparticles for gene delivery has also been studied. In gene delivery,<br />

tissue targeting is very important for the efficacy and safety of treatment. To date, the transfection<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 223<br />

efficiency offered <strong>by</strong> synthetic gene delivery systems is very low compared to that of viral vectors.<br />

This is also true for nanoparticle-based systems. Extensive research is being done in the area to<br />

improve transfection efficiencies <strong>by</strong> designing better delivery systems with improved targeting<br />

ability, DNA stabilization, and cellular interaction.<br />

Mansouri reported on the formulation of nanoparticles through the complex coacervation of<br />

chitosan-folic acid conjugates with DNA. 69 Chitosan is a biocompatible polycation that joins to<br />

form a complex with DNA through electrostatic interactions and provides protection against<br />

nuclease degradation. Complex coacervation is an optimal preparation technique for the encapsulation<br />

of DNA because it avoids the use of organic solvents and high-energy ultrasonication.<br />

Results revealed that the integrity of plasmid DNA in the particles was maintained and that<br />

conjugation of chitosan to folic acid did not interfere with the electrostatic interaction between<br />

chitosan and DNA. As expected, the ratio of chitosan to DNA, and consequent charge ratio, had an<br />

effect on particle size and zeta potential. Nanoparticles smaller than 200 nm were obtained with a<br />

chitosan amino group to DNA phosphate group ratio of more than 2. The use of these nanoparticles<br />

consequently offers a promising alternative for nonviral gene therapy for the treatment of cancer<br />

and other diseases in which folate receptors are overexpressed.<br />

Nanoparticles with a poly(L-lactic acid) (PLL) core and a polyethyleneimide (PEI) surface<br />

conjugated to folate were utilized for delivery of plasmid DNA. 70 Folic acid was conjugated to<br />

the N-terminal amino group of PEI. PEI is a polycation that has been used in the past for DNA<br />

condensation and delivery because it protects DNA from degradation through an endosomal escape<br />

mechanism. Here nanoparticles were prepared through the self-assembly of the amphiphilic folate-<br />

PEI–PLL copolymer with DNA in an aqueous medium. Nanoparticles of approximately 100–<br />

150 nm in diameter and spherical shape were produced. In vitro luciferase transfection studies<br />

revealed that this system actually resulted in lower luciferase expression than<br />

PEI–DNA complexes.<br />

In a separate report, folate-polyethyleneglycol–distearoylphophatidylethanolamine conjugate<br />

(f-PEG–DSPE), 3([N-(N 0 ,N 0 -dimethylaminoethane)-carbamoyl] cholesterol, and Tween 80 were<br />

used to complex with DNA into cationic nanoparticles of 100–200 nm in diameter with a modified<br />

ethanol injection method. 71 The formulation was carried out <strong>by</strong> dissolving the lipids in ethanol and<br />

then removing the solvent through evaporation in the presence of water. The folate moiety, which is<br />

conjugated to the PEG end of one of the lipid conjugates, naturally localizes at the surface of the<br />

nanoparticles because of PEG migration toward the water phase. Tween 80, a nonionic surfactant,<br />

and PEG were incorporated with the purpose of improving the in vivo stability of the cationic<br />

nanoparticles through steric hindrance. The size of the nanoparticles with higher PEG content was<br />

maintained in the presence of serum. This suggests that these nanoparticles are better able to<br />

maintain their structural integrity in the presence of anionic competitors present in blood. Folate<br />

targeting enhanced association and transfection efficacy of nanoparticles complexed with a luciferase-encoding<br />

plasmid on FR(C) KB cells. The association and efficacy were reduced when folic<br />

acid was present in the medium, thus revealing the involvement of the folate receptor in the<br />

transport of the plasmid DNA into the cells.<br />

A possible limitation of folate targeting is the noted variability of FR expression levels not only<br />

between patients, but also within a single tumor. 44 In addition, it has been reported that expression<br />

of FR in cancerous cell lines is not representative of those one sees in vivo. 44 Consequently,<br />

screening protocols for FR expression will need to be utilized clinically in order to determine if<br />

folate-targeted therapies are appropriate.<br />

12.6 APPROACHES FOR CANCER TARGETING TO SPECIFIC CANCER TYPES<br />

In a recent review, Kim and Nie describe passive and active targeting methods and then go into<br />

detail on a number of active targeting techniques. 72 The active target combinations they mention<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


224<br />

are lectin–carbohydrate, ligand–receptor and antibody–antigen. One limitation of using these<br />

targeting strategies is that the lectin–carbohydrate targeting systems are usually targeted to<br />

whole organs, making them inappropriate for targeting a cancerous part of a particular organ or<br />

tissue. New antibody systems show a great deal of promise. Unfortunately, they also have potentially<br />

harmful side effects such as advanced gastric adenocarcinoma. The latter arises when one<br />

attempts to target breast cancer on account of the fact that antigen-positive normal cells to the<br />

antibody BR96 in gastric mucosa, small intestine, and pancreas. Some of the aspects of angiogenic<br />

targeting mentioned here have already been covered elsewhere in this chapter. Specific targeting<br />

systems that are in use as cancer therapeutics are shown in Table 12.1. 8,72<br />

12.6.1 PROSTATE CANCER<br />

Recently, aptamers have been used to target nanoparticulate systems to prostate-specific membrane<br />

antigen, a known prostate cancer tumor marker. A model drug, rhodamine-labeled dextran, was<br />

encapsulated in PEGylated poly(lactic acid) nanoparticles, which were subsequently surface<br />

modified with a prostate specific RNA aptamer (A10). Binding of the aptamer nanoparticles to<br />

LNCaP cells expressing prostate specific membrane antigen in vitro was significantly enhanced<br />

when compared to a control of nontargeted particles. Additionally, very low binding was seen on<br />

nonprostate specific membrane antigen expressing cells (PC3). The nanoparticles were shown to<br />

both target and be taken up <strong>by</strong> the prostate cancer epithelial cells. This evidence points to the<br />

conclusion that this novel aptamer-based, targeted nanoparticle delivery approach can be effective. 73<br />

Gao and collaborators report the use of quantum dots (QD) for in vivo targeting of prostate cancer<br />

and imaging of tumors. 7 The core-shell CdSe–ZnS quantum dots contain tri-n-octylphosphine<br />

oxide (TOPO) that binds to a covering of high molecular weight ABC triblock copolymer of<br />

polybutylacrylate, polyethylacrylate, polymethacrylic acid, and an 8-carbon alkyl side chain.<br />

The complex is functionalized with PEG molecules and monoclonal antibodies to prostate-specific<br />

membrane antigen. Specific binding was shown for prostate cancer lines whereas low binding was<br />

seen to normal cells. The QD-antibody formulations were studied in vivo in a mouse model human<br />

prostate cancer. The nanoparticles were shown to target to the tumor both <strong>by</strong> passive and active<br />

antibody targeting. Sensitive and multicolor fluorescence imaging of cancer cells in vivo was<br />

TABLE 12.1<br />

Targeting Systems Utilizing Antibodies Currently in Use to Treat Cancer<br />

Mechanism<br />

Antibody<br />

Target Trade Name<br />

Agonist activity CD40, CD137 Various<br />

Antagonist activity CTLA4 MDX-010<br />

Angiogenesis inhibition VEGF Avastine<br />

Antibody-dependent cell-mediated cytotoxicity CD20 Rituxan w , HuMax-CD20, Zevalin w<br />

Inhibition of binding of extracellular growth signals HER-2/neu Herceptin<br />

Receptor blockage EGF receptor HuMax-EGFr<br />

Toxin-mediated killing CD33 Mytotarg w<br />

Disruption signaling HER-2/neu Pertuzumab (2C4)<br />

Complement-dependent cytotoxicity CD20 Rituxan w , HuMax-CD20<br />

Blockage ligand binding EGF receptor Erbutixe<br />

Antibody-dependent lysis of leukemic cells following<br />

cell binding<br />

CD52 Campath w<br />

Inhibits phosphorilation of tyrosine kinases EGF receptor Iressa<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 225<br />

accomplished in this study. Nonspecific uptake was seen in the liver and spleen with little or no<br />

uptake in other organs.<br />

12.7 TARGETED NANOPARTICLES AND IMAGING OF CANCER<br />

Hofmann and collaborators have evaluated the ability of super paramagnetic iron oxide nanoparticles<br />

(SPION) to interact with human melanoma cells in such a way that these particles could be<br />

selectively targeted to tumor cells ands then imaged using MRI. 74 They varied the coating placed on<br />

these particles and found that when comparing particles coated with poly(vinyl alcohol) (PVA), a<br />

vinyl alcohol/vinyl amine copolymer (amine-SPION), PVA with randomly distributed carboxylic<br />

groups or PVA with randomly distributed thiol groups, human cells in culture would interact strongly<br />

only with the amine-SPION particles. Furthermore, these particles showed the lowest cytotoxicity.<br />

This human study involved intravenous administration of Combidex (Advanced Magnetics,<br />

Inc., ferumoxtran-10, a molecular imaging agent of iron oxide nanoparticles and a dense packing of<br />

dextran derivatives) to 18 men ages 21–46 with diagnosed testicular cancer. 75 The Combidex was<br />

imaged using MRI. From this study, it seems evident that those lymph nodes with a higher signal<br />

were classified as being malignant. However, based on the information from Advanced Magnetics,<br />

these particles should accumulate selectively in noncancerous lymph node tissue. The particles are<br />

still experimental and rightly so.<br />

Although treatment of cancer with targeted nanoparticles is an important goal, more accurate<br />

imaging of cancer is needed to allow for the optimal treatment for each patient. Towards that end,<br />

a considerable amount of research is underway with various imaging techniques to establish more<br />

accurate determination of the presence and extent of cancer growth and metastases. Because magnetic<br />

resonance imaging (MRI) is a widely used imaging technique, much work is currently being done to<br />

develop targeted imaging agents for MRI. Some of these involve paramagnetic and superparamagnetic<br />

iron oxides due to their ability to affect water relaxation times T1 and T2. Gasco and collaborators<br />

have prepared solid lipid nanoparticles containing Endorem, superparamagnetic iron oxide nanoparticles<br />

(Guebert and Advanced Magnetics), using either a multiple emulsion technique or an oil in<br />

water emulsion technique. 76 Although the loading rates achieved were less than 1 wt% iron, it was<br />

possible to detect and image in vivo in rats. Incorporation of the Endorem in the SLN allowed passage<br />

across the blood brain barrier, passage which was not possible with Endorem alone.<br />

Often development of nanoparticle systems is a two-pronged approach involving both drug and<br />

imaging agent. If a targeting system is successful, one should be able to enhance the imaging of a<br />

cancer and then kill it with the same system. One such study involved the preparation of glycol–<br />

chitosan nanoaggregates to which either fluorescein isothiocyanate (FITC) or doxorubicin (Dox)<br />

was conjugated. 12 Based on a single tail-vein injection of FITC-conjugates, levels remained high for<br />

eight days and gradually increased in the tumors of rats with II45 mesothelioma cells, in the kidney,<br />

and to a lesser extent in the spleen. Meanwhile, levels decreased in other organs. The liver showed<br />

some accumulation at day 3 but was significantly lower at day 8 relative to days 1 and 3. The<br />

performance of these systems, as evidenced <strong>by</strong> a decrease in tumor volume, was excellent with a<br />

consistent decrease in tumor volume after day 13 when a tail-vein injection of Dox-nanoaggregates<br />

is given at days 13 and 19.<br />

12.8 OTHER TARGETS FOR CANCER<br />

A new approach to targeting is the use of lectins, which are plant proteins that specifically recognize<br />

cell surface carbohydrates. The latter function as selective cancer-cell-targeting agents. PLGA<br />

nanoparticles of mean diameter 331 nm incorporating isopropyl myristate were used to deliver<br />

paclitaxel to malignant A549 and H1299 and normal CCL-186 pulmonary cells in vitro <strong>by</strong> means of<br />

wheat germ agglutinin lectin as the targeting molecule. The in vitro cytotoxicity against A549 and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


226<br />

H1299 cells was significantly increased with wheat germ agglutinin-conjugated PLGA nanoparticles<br />

containing IPM and paclitaxel compared to controls. 77 In a subsequent study, Mo and Lim<br />

evaluated in vivo efficacy of the same nanoparticles in a SCID mouse model injected with an A549<br />

tumor nodule. 78 One injection of wheat germ agglutinin-conjugated PLGA nanoparticles<br />

containing IPM and a paclitaxel dose of 10 mg/kg inhibited tumor growth without appreciable<br />

weight loss. Tumor doubling was increased to 25 days compared to 11 days for conventionally<br />

formulated paclitaxel.<br />

12.9 AVIDIN AND BIOTIN TARGETING<br />

Although it is <strong>by</strong> far the most widely utilized polymer for surface modification of nanoparticles,<br />

PEG is not the only compound that can be included at the surface of nanoparticles. Nor is it<br />

necessary to achieve active targeting. Saltzman has recently reported a method for incorporating<br />

avidin-fatty acid conjugates into the surface of PLGA nanoparticles. 79 This method resulted in<br />

avidin at the surface of the nanoparticles that remained active for weeks. The ability of these<br />

nanoparticles to target to biotin was verified <strong>by</strong> targeting of the nanoparticles to biotinylated<br />

agarose beads. Not only could this system be used for targeted delivery; it can also be utilized to<br />

selectively modify surfaces for tissue engineering.<br />

In another such example, Hunziker has prepared biotin-functionalized (poly(2-methyloxazoline)-b-poly(dimethylsiloxane)-b-poly(2-methyl-oxazoline)<br />

triblock copolymers. 80 The<br />

biotinylated targeting agents were added using streptavidin as a coupling agent. Uptake of these<br />

“nanocontainers” was seen in the presence of the target receptor, the macrophage scavenger<br />

receptor SRA1, but not in the absence of this target.<br />

12.10 CONCLUSIONS<br />

The amount of research in targeted, polymeric nanoparticles for cancer imaging and therapy has<br />

increased dramatically in the past 5–10 years. Seeing actual products using targeted therapies has<br />

no doubt fueled that work. In the next decade, we will certainly see products, whether with<br />

polymeric nanoparticles or some other type of delivery system, using folate receptors and carrying<br />

imaging agents. All of these technologies, driven <strong>by</strong> the fields of fundamental immunology,<br />

biochemistry, polymer chemistry, and biomedical engineering, are bringing us closer to the time<br />

when cancer may be treated on an individual basis. One patient’s diagnosis and treatment will be<br />

unique to her condition and will be the most effective treatment possible for her. Until other<br />

scientists determine how to stop cancer from occurring, those mentioned in this chapter and<br />

many more besides them are doing their best to eliminate cancer.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Ferlay, J. et al., GLOBOCAN 2002: Cancer Incidence, Mortality and Prevalence Worldwide, IARC<br />

CancerBase No. 5, Version 2.0., IARC Press, Lyon, 2004.<br />

2. Society, A. C., Cancer Facts and Figures 2005, American Cancer Society, Atlanta, 2005.<br />

3. Yokoyama, M. and Okano, T., Targetable drug carriers: Present status and a future perspective,<br />

Advanced Drug Delivery Reviews, 21, 77–80, 1996.<br />

4. Torchilin, V. P., Strategies and means for drug targeting: An overview, In Biomedical Aspects of Drug<br />

Targeting, Muzykantov, V. and Torchilin, V., Eds., Kluwer Academic, Norwell, pp. 3–26, 2002.<br />

5. McDonald, D. M. and Baluk, P., Significance of blood vessel leakiness in cancer, Cancer Research,<br />

62, 5381–5385, 2002.<br />

6. Conner, S. D. and Schmid, S. L., Regulated portals of entry into the cell, Nature, 422, 37–44, 2003.<br />

7. Gao, X. et al., In vivo cancer targeting and imaging with semiconductor quantum dots, Nature<br />

Biotechnology, 22(8), 969–976, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 227<br />

8. Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancer therapy,<br />

Advanced Drug Delivery Reviews, 56, 1649–1659, 2004.<br />

9. Luo, Y. et al., Targeted delivery of doxorubicin <strong>by</strong> HPMA copolymer-hyaluronan bioconjugates,<br />

Pharmaceutical Research, 19, 396–402, 2002.<br />

10. Mastrobattista, E. et al., Targeted liposomes for delivery of protein-based drugs into the cytoplasm of<br />

tumor cells, Journal of Liposome Research, 12, 57–65, 2002.<br />

11. Yoo, H. S. and Park, T. G., Folate receptor targeted biodegradable polymeric doxorubicin micelles,<br />

Journal of Controlled Release, 96, 273–283, 2004.<br />

12. Vauthier, C. et al., Drug delivery to resistant tumors: The potential of poly(alkly cyanoacrylate)<br />

nanoparticles, Journal of Controlled Release, 93, 151–160, 2003.<br />

13. Peracchia, M. T., Stealth nanoparticles for intravenous administration, S.T.P. Pharma, 13(3),<br />

155–161, 2003.<br />

14. Shenoy, D. B. and Amiji, M. M., Poly(ethylene oxide)-modified poly(e-caprolactone) nanoparticles<br />

for targeted delivery of tamoxifen in breast cancer, International Jounal of Pharmaceutics, 293,<br />

261–270, 2005.<br />

15. Avgoutsakis, K. et al., PLGA–mPEG nanoparticles of cisplatin: In vitro nanoparticles degradation,<br />

in vitro drug release and in vivo drug residence in blood properties, Journal of Controlled Release, 79,<br />

125–135, 2002.<br />

16. Machida, Y. et al., Efficacy of nanoparticles containing irinotecan prepared using poly(DL-lactic acid)<br />

and poly(ethylene glycol)-poly(propylene glycol)-poly(ethylene glycol) against M5076 tomur in the<br />

early liver metastatic stage, S.T.P. Pharma, 13(4), 225–230, 2003.<br />

17. Brigger, I. et al., Poly(ethylene glycol)-coated hexadecylcyanoacrylate nanospheres display a<br />

combined effect for brain tumor targeting, Journal of Pharmacology and Experimental Therapeutics,<br />

303, 928–936, 2002.<br />

18. Reddy, L. H., Sharma, R. K., and Murthy, R. S. R., Enhanced tumour uptake of doxorubicin loaded<br />

poly(butyl cyanoacrylate) nanoparticles in mice bearing Dalton’s lymphoma tumour, Journal of Drug<br />

Targeting, 12(7), 443–451, 2004.<br />

19. Kaul, G. and Amiji, M., Biodistribution and targeting potential of poly(ethylene glycol)-modified<br />

gelatin naoparticlea in subcutaneous murine tumor model, Journal of Drug Targeting, 12(9–10),<br />

585–591, 2004.<br />

20. Folkman, J., Angiogenic-dependent diseases, Seminars in Oncology, 28(6), 536–542, 2001.<br />

21. Lyden, D. et al., Impaired recruitment of bone-marrow-derived endothelial and hematopoietic<br />

presursor cells blocks tumor angiogenesis and growth, Nature Medicine, 7(11), 1194–1201, 2001.<br />

22. Folkman, J., Vascularization of tumors, Scientific American, 234(5), 59–76, 1976.<br />

23. Folkman, J., Fundamental concepts of the angiogenic process, Current Molecular Medicine, 3,<br />

643–651, 2003.<br />

24. Hanahan, D. and Folkman, J., Patterns and emerging mechanisms of the angiogenic switch during<br />

tumorgenesis, Cell, 86, 353–364, 1996.<br />

25. Carmeliet, P. and Jain, R. K., Angiogenesis in cancer and other diseases, Nature, 407, 249–257, 2000.<br />

26. Leenders, W. P. J., Targetting VEGF in anti-angiogenic and anti-tumour therapy: Where are we now?,<br />

International Journal of Experimental Pathology, 79, 339–346, 1998.<br />

27. Claffey, K. P. et al., Expression of vascular permeasbility factor/vascular endothelial growth factor <strong>by</strong><br />

melanoma cells increases tumor growth, angiogenesis, and experimental metastasis, Cancer Research,<br />

56, 172–181, 1996.<br />

28. Potgens, A. J. G. et al., Analysis of the tumor vasculature and metastatic behavior of xenografts of<br />

human melanoma cells lines transfected with vascular permeability factor, American Journal of<br />

Pathology, 148(4), 1203–1217, 1996.<br />

29. Li, L. et al., A novel antiangiogenesis therapy using an integrin antagonist or anti-FLK-1 antibody<br />

coated with 90 Y-labeled nanoparticles, International Journal of Radiation Oncology Biology Physics,<br />

58(4), 1215–1227, 2004.<br />

30. Brooks, P. C., Clark, R. A. F., and Cheresh, D. A., Requirement of vascular integrin a vb 3 for<br />

angiogenesis, Science, 264, 569–571, 1994.<br />

31. Varner, J. A. and Cheresh, D. A., Integrins and cancer, Current Opinion in Cell Biology, 8, 724–730,<br />

1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


228<br />

Nanotechnology for Cancer Therapy<br />

32. Hood, J. D. et al., Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science, 296,<br />

2404–2407, 2002.<br />

33. Schiffelers, R. M. et al., Cancer siRNA therapy <strong>by</strong> tumor selective delivery with ligand-targeted<br />

sterically stabilized nanoparticle, Nucleic Acids Research, 32(19), e149, 2004.<br />

34. Kopelman, R. et al., Multifunctional nanoparticle platforms for in vivo MRI enhancement and photodynamic<br />

therapy of a rat brain cancer, Journal of Magnetism and Magnetic Materials, 293, 404–410,<br />

2005.<br />

35. Bib<strong>by</strong>, B. C. et al., Pharmacokinetics and biodistribution of RGD-targeted doxorubicin-loaded nanoparticles<br />

in tumor-bearing mice, International Jounal of Pharmaceutics, 293, 281–290, 2005.<br />

36. Mulder, W. J. M. et al., MR molecular imaging and fluorescence microscopy for identification of<br />

activated tumor endothelium using a bimodal lipidic nanoparticle, The FASEB Journal, 19,<br />

2008–2010, 2005.<br />

37. Lanza, G. M., Magnetic resonance molecular imaging and targeted drug delivery with site-specific<br />

nanoparticles. In BioMEMS and Biomedical Nanotech World, Columbus, OH, 2002.<br />

38. Leamon, C. P. et al., Synthesis and biological evaluation of EC72: A new folate-targeted chemotherapeutic,<br />

Bioconjugate Chemistry, 16, 803–811, 2005.<br />

39. Lee, J. W. et al., Synthesis and evaluation of taxol-folic acid conjugates as targeted antineoplastics,<br />

Bioorganic Medicinal Chemistry, 10, 2397–2414, 2002.<br />

40. Paranjpe, P. V. et al., Tumor-targeted bioconjugate based delivery of camptothecin: Design, sythesis<br />

and in vitro evaluation, Journal of Controlled Release, 100, 275–292, 2004.<br />

41. Gabizon, A. A., Liposome circulation time and tumor targeting: Implications for cancer<br />

chemotherapy, Advanced Drug Delivery Reviews, 16, 285–294, 1995.<br />

42. Lee, D. H., Kim, D. I., and Bae, Y. H., Doxorubicin loaded pH-sensitive micelle targeting acidic<br />

extracellular pH of human ovarian A2780 tumor in mice, Journal of Drug Targeting., 13(7), 391–397,<br />

2005.<br />

43. Douglas, J. T. et al., Targeted gene delivery <strong>by</strong> tropism-modified adenoviral vectors, Nature Biotechnology,<br />

14, 1574–1578, 1996.<br />

44. Elnakat, H. and Ratnam, M., Distribution, functionality and gene regulatio of folate receptor isoforms:<br />

Implications in targeted therapy, Advanced Drug Delivery Reviews, 56, 1067–1084, 2004.<br />

45. Leamon, C. P. and Reddy, J. A., Folate-targeted chemotherapy, Advanced Drug Delivery Reviews, 56,<br />

1127–1141, 2004.<br />

46. Sudimack, J. and Lee, R. J., Targeted drug delivery via the folate receptor, Advanced Drug Delivery<br />

Reviews, 41, 147–162, 2000.<br />

47. Antony, A., The biological chemistry of folate receptors, Blood, 79, 2807–2820, 1992.<br />

48. Kamen, B. and Capdevila, A., Receptor-mediated folate accumulation is regulated <strong>by</strong> the cellular<br />

folate content, Proceedings of the National Academy of Sciences of the United State of America, 88,<br />

5983–5987, 1986.<br />

49. Mattes, M. et al., Patterns of antigen distribution in human carcinomas, Cancer Research, 50,<br />

880s–884s, 1990.<br />

50. Coney, L. et al., Distribution of the folate receptor GP38 in normal and malignant cell lines and<br />

tissues, Cancer Research, 52, 3396–3401, 1991.<br />

51. Weitman, S. et al., Distribution of the folate receptor GP38 in normal and malignant cell lines and<br />

tissues, Cancer Research, 52, 3396–3401, 1992.<br />

52. Ross, J., Chaadhuri, P., and Ratnam, M., Differential regulation of folate receptor isoforms in normal<br />

and malignant cell lines. Physiologic and clinical implications, Cancer, 73, 2432–2443, 1994.<br />

53. Weitman, S. et al., Cellular localization of thr folate receptor: Potential role in drig toxicity and folate<br />

homeostasis, Cancer Research, 52, 6708–6711, 1992.<br />

54. Weitman, S., Frazier, K., and Kamen, B., The folate receptor in central nervous system malignancies<br />

of childhood, Journal of Neurooncology, 21, 107–112, 1994.<br />

55. Garin-Chesa, P. et al., Trophoblast and ovarian cancer antigen LK26. Sensitivity and specificity in<br />

immunopathology and molecular identification as a folate-binding protein, American Journal of<br />

Pathology, 142, 557–567, 1993.<br />

56. Toffoli, G. et al., Overexpression of folate binding protein in ovaian cancers, International Journal of<br />

Cancer, 74, 193–198, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Nanoparticles for Tumor-Targeted Drug Delivery 229<br />

57. Holm, J. et al., High-affinity folate binding in human choroid plexus. Characterization of radioligand<br />

binding, immunoreactivity, molecular heterogeneity and hydrophobic domain of the binding protein,<br />

Biochemical Journal, 280, 267–271, 1991.<br />

58. Shen, F. et al., Identification of a novel folate receptor, a truncated receptor and receptor type B in<br />

hematopoietic cells: cDNA cloning, expression, immunoreactivity, and tissue specificity, Biochemistry,<br />

33, 1209–1215, 1994.<br />

59. Sadasivan, E. et al., Characterization of multiple forms of folate-binding protein from himan leukemia<br />

cells, Biochimica et Biophysica Acta, 882, 311–321, 1986.<br />

60. Sadasivan, E. et al., Purification, properties, and immunological characterization of folate-binding<br />

proteins from human leukemia cells, Biochimica et Biophysica Acta, 925, 36–47, 1987.<br />

61. Morikawa, J. et al., Identification of signature genes <strong>by</strong> microarray for acute myeloid leukemia<br />

without maturation and acute promyelocytic leukemia with t(15;17)(q22;q12)(PML/RARalpha),<br />

International Journal of Oncology, 23, 617–625, 2003.<br />

62. Hilgenbrink, A. R. and Low, P. S., Folate receptor-mediated drug targeting: From therapeutics to<br />

diagnostics, Journal of Pharmaceutical Sciences, 94(10), 2135–2146, 2005.<br />

63. Sabharanjak, S. and Mayor, S., Folate receptor endocytosis and trafficking, Advanced Drug Delivery<br />

Reviews, 56, 1099–1109, 2004.<br />

64. Coliva, A. et al., 90Y Labeling of monoclonal antibody MOv18 and preclinical validation for radioimmunotherapy<br />

of human ovarian carcinomas, Cancer Immunology, Immunotherapy, 54, 1200–1213,<br />

2005.<br />

65. Park, E. K., Lee, S. B., and Lee, Y. M., Preparation and characterization of methoxy poly(ethylene<br />

glycol)/poly(3-caprolactone) amphiphilic block copolymeric nanospheres for tumor-specific folatemediated<br />

targeting of anticancer drugs, Biomaterials, 26, 1053–1061, 2005.<br />

66. Kim, S. H. et al., Target-specific cellular uptake of PLGA nanoparticles coated with poly(L-lysine)poly(ethylene<br />

glycol)-folate conjugate, Langmuir, 21, 8852–8857, 2005.<br />

67. Kukowska-Latallo, J. F. et al., Nanoparticle targeting of anticancer drug improves therapeutic<br />

response in animal model of human epithelial cancer, Cancer Research, 65(12), 5317–5324, 2005.<br />

68. Nayak, S. et al., Folate-mediated cell targeting and cytotoxicity using thermoresponsive microgels,<br />

Journal of the American Chemistry Society, 126, 10258–10259, 2004.<br />

69. Mansouri, S. et al., Characterization of folate–chitosan–DNA nanoparticles for gene therapy, Biomaterials,<br />

27(9), 2060–2065, <strong>2006</strong>.<br />

70. Wang, C.-H. and Hsiue, G.-H., Polymer-DNA hybrid nanoparticles based on folate-polyethyleneimine-block-poly(L-lactide),<br />

Bioconjugate Chemistry, 16, 391–396, 2005.<br />

71. Hattori, Y. and Maitani, Y., Enhanced in vitro DNA transfection efficiency <strong>by</strong> novel folate-linked<br />

nanoparticles in human prostate cancer and oral cancer, Journal of Controlled Release, 97, 173–183,<br />

2004.<br />

72. Kim, G. J. and Nie, S., Targeted cancer nanotherapy, Nanotoday, August, 28–33, 2005.<br />

73. Farokhzad, O. C. et al., Nanoparticles-aptamer bioconjugates: A new approach for targeting prostate<br />

cancer cells, Cancer Research, 64, 7668–7672, 2004.<br />

74. Petri-Fink, A. et al., Development of functionalized superparamagnetic iron oxide nanoparticles for<br />

interaction with human cancel cells, Biomaterials, 26, 2685–2694, 2005.<br />

75. Harisinghani, M. G. et al., A pilot study of lymphotrophic nanoparticle-enhanced magnetic resonance<br />

imaging technique in early stage testicular cancer: A new method for noninvasive lymph node<br />

evaluation, Urology, 66, 1066–1071, 2005.<br />

76. Peira, E. et al., In vitro and in vivo study of solid lipid nanoparticles loaded with superparamagnetic<br />

iron oxide, Journal of Drug Targeting, 11(1), 19–24, 2003.<br />

77. Mo, Y. and Lim, L.-Y., Preparation and in vitro anticancer activity of wheat germ agglutinin (WGA)conjugated<br />

PLGA nanoparticles loaded with paclitaxel and isopropyl myristate, Journal of Controlled<br />

Release, 107, 30–42, 2005.<br />

78. Mo, Y. and Lim, L., Paclitaxel-loaded PLGA nanoparticles: Potentiation of anticancer activity <strong>by</strong><br />

surface conjugation with wheat germ agglutinin, Journal of Controlled Release, 108, 244–262, 2005.<br />

79. Fahmy, T. M. et al., Surface modification of biodegradable polyesters with fatty acid conjugates for<br />

improved drug targeting, Biomaterials, 26, 5727–5736, 2005.<br />

80. Broz, P. et al., Cell targeting <strong>by</strong> a generic receptor-targeted polymer nanocontainer platform, Journal<br />

of Controlled Release, 102, 475–488, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


13<br />

CONTENTS<br />

Long-Circulating Polymeric<br />

Nanoparticles for Drug and<br />

Gene Delivery to Tumors<br />

Sushma Kommareddy, Dinesh B. Shenoy,<br />

and Mansoor M. Amiji<br />

13.1 Introduction ....................................................................................................................... 231<br />

13.2 Polymeric Nanoparticles ................................................................................................... 232<br />

13.3 Nanoparticles for Drug Delivery ...................................................................................... 233<br />

13.4 Nanoparticles for Gene Delivery ...................................................................................... 233<br />

13.5 Long-Circulating Nanoparticles ........................................................................................ 234<br />

13.6 Illustrative Examples of Polymeric Nanoparticles for Drug Delivery............................. 235<br />

13.6.1 Polyethylene Oxide-Modified Poly(b-Amino Ester) Nanoparticles .................. 235<br />

13.6.2 Poly(Ethylene Oxide)-Modified Poly(3-Caprolactone) Nanoparticles............... 236<br />

13.7 Illustrative Examples of Polymeric Nanoparticles for Gene Delivery ............................ 237<br />

13.7.1 Polyethylene Glycol-Modified Gelatin Nanoparticles........................................ 237<br />

13.7.2 Polyethylene Glycol-Modified Thiolated-Gelatin Nanoparticles ....................... 238<br />

13.8 Concluding Remarks ......................................................................................................... 238<br />

Acknowledgments ........................................................................................................................ 239<br />

References .................................................................................................................................... 239<br />

13.1 INTRODUCTION<br />

During the past few decades, there have been extensive efforts in the treatment and cure of cancer<br />

that is still the second leading cause of death next to the cardiovascular diseases. With advances in<br />

research techniques, there has been a surge of therapeutic agents in the form of proteins and nucleic<br />

acids as a source of new chemical entities for the treatment of cancer. 1–7 However, effective<br />

delivery of these novel agents to the target tissue has always been a problem.<br />

The therapeutic agents used in cancer treatment are generally administered in the systemic<br />

circulation. The drug carrier, therefore, must overcome physiological barriers to reach the tumor<br />

cell in sufficient concentrations and to reside for the necessary duration to exert the pharmacological<br />

effect. These barriers include transport of the drugs within the blood vasculature and<br />

transport from the vasculature into the surrounding tumor tissues and through the interstitial spaces<br />

within the tumor. Solid tumors are characterized <strong>by</strong> vasculature that is heterogeneous in size and<br />

distribution, having a central avascular/necrotic region and vascularized peripheral region with<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

231


232<br />

discontinuous endothelium in the microvessels. Depending on the anatomic region of the tumor, the<br />

pore size of the endothelial junctions is found to vary from 100 to 780 nm with a mean of<br />

approximately 400 nm. 8–11 Tumor vasculature is also characterized <strong>by</strong> a lack of lymphatic drainage.<br />

12–14 In addition to the complexity of tumor physiology, the development of multi-drug<br />

resistance (MDR) in tumor cells exacerbates the problem of achieving selective toxicity in<br />

tumor tissues.<br />

Typically, a tumor consists of neoplastic cells, stromal cells, and the extracellular matrix and is<br />

associated with blood vessels, nerves, and immune cells that form an integral part of the stroma. 15,16<br />

The tumor also contains leukocytes, fibroblasts, and other extracellular elements. The majority of<br />

the infiltrated leukocytes are macrophages, more popularly referred to as tumor-associated macrophages<br />

(TAM). They have increased phagocytic activity and play an active role in tumor<br />

progression. The presence of these activated macrophages in large numbers might lead to enhanced<br />

clearance of a drug-carrying vector from the tumor microenvironment. 15–18<br />

13.2 POLYMERIC NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

Sub-micronic particles prepared from polymers are becoming increasingly popular for the delivery<br />

of drugs or genes to tumor tissues. When compared to all other colloidal delivery systems, nanoparticles<br />

have better stability in plasma and higher encapsulation efficiency, and they are amenable<br />

to large-scale preparation. These nanoparticles can be used to increase the solubility of hydrophobic<br />

drugs, lower the toxicity of drugs with a high therapeutic index, and enhance the stability of the<br />

payload, in the case of DNA, <strong>by</strong> protecting it from degradation <strong>by</strong> extracellular enzymes. Furthermore,<br />

these particles permit controlled release of the payload at the target site at relatively low<br />

doses. 19 In addition, there is a wide range of polymeric materials whose physicochemical properties<br />

can be tailored to achieve polymeric nanoparticles of the required nature.<br />

Passive targeting of these particles at the tumor sites is generally achieved <strong>by</strong> the enhanced<br />

vascular permeability and lack of lymphatic drainage, together termed the enhanced permeability<br />

and retention effect (EPR) (Figure 13.1). 12–14 The accumulation of polymeric nanoparticles<br />

carrying drugs is dependent on the physical chemistry of the polymer, including molecular<br />

weight, surface charge, nature of the polymer, etc. The EPR is one of the main reasons for the<br />

success of polymeric nanoparticles in targeting tumors. Site-specific or active targeting of these<br />

particles can be achieved <strong>by</strong> conjugating with targeting moieties or ligands that are specific to the<br />

tumor cells. These targeting moieties present on the polymer backbone are used to exploit<br />

the differences between tumor cells and normal cells through receptor-mediated endocytosis.<br />

Active targeting, in particular, can be used to overcome the obstacle of tumor metastasis and to<br />

increase the specificity and efficacy of the polymeric carrier system. Transferrin, folate, epidermal<br />

growth factor, and argenine–glycine–aspartic acid (RGD) tripeptide are some of the moieties that<br />

are used for targeting tumor cells. 7,20–26<br />

The major drawback to the use of nanoparticles in cancer therapy is their initial burst release of<br />

the drug upon administration into systemic circulation that is followed <strong>by</strong> slow controlled release<br />

of the encapsulated drug. The polymers used in the preparation of nanoparticles generally lack the<br />

ability to encapsulate both hydrophilic and hydrophobic drugs in the same polymer system (i.e., the<br />

hydrophilic drugs have to be encapsulated in a hydrophilic polymer). Some of the synthetic polymers<br />

have cytotoxicity issues associated with them. However, this difficulty could be overcome <strong>by</strong><br />

using biosynthetic polymers that are biodegradable and modified to have physicochemical properties<br />

similar to those of the synthetic polymers. The use of polymeric nanoparticles for systemic<br />

delivery must be limited to non-cationic or surface-modified polymers to prevent adsorption of<br />

plasma proteins onto the surface of the nanoparticles. Furthermore, some of the naturally occurring<br />

polymers have immunogenic reactions when injected into blood; such reactions could be prevented<br />

through the use of non-immunogenic biopolymers now available.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Long-Circulating Polymeric Nanoparticles for Drug and Gene Delivery to Tumors 233<br />

Extravascular Space<br />

PEG-Modified Polymeric Nanoparticles<br />

Blood flow<br />

Extravascular Space<br />

13.3 NANOPARTICLES FOR DRUG DELIVERY<br />

The majority of the chemotherapeutic agents used in cancer therapy are low molecular weight drugs<br />

with a high volume of distribution, leading to the presence of toxic compounds throughout the<br />

body. These drugs must be given in high doses in order to reach therapeutic levels in tumor tissues.<br />

In the process, healthy tissues are exposed to cytotoxic drugs, resulting in side effects such as bone<br />

marrow suppression, alopecia, anorexia, etc. The low molecular weight anti-cancer agents are also<br />

rapidly cleared from the body. Continuous treatment of tumor cells with these drugs would exacerbate<br />

the problem of MDR that is one of the main reasons for the failure of chemotherapy. 19<br />

Numerous efforts in improving cancer chemotherapy have focused on increasing the therapeutic<br />

index and reducing the non-selective cytotoxicity of the drugs in use. Macromolecular<br />

carriers such as polymer conjugates, liposomes, microspheres, and nanoparticles are used for<br />

this purpose. Of these, the nanoparticles have been promising, mainly because of their advantages<br />

over other novel drug delivery vehicles such as their safety, enhanced stability, possibility of<br />

tailoring and surface modification, and industrial scale-up. 27 Nanoparticles prepared from<br />

polymer–drug conjugates have altered pharmacokinetic distribution and increased pharmacological<br />

activity at tolerable doses. This is due mainly to the controlled release of the drug and reduced renal<br />

clearance of the low molecular weight drugs.<br />

The concept of targeting drugs at the site of action was first described as “magic bullets” <strong>by</strong><br />

Paul Ehrlich. 28 Later, Ringsdorf presented a model of the polymer–drug conjugate that could be<br />

used to improve chemotherapy drugs for cancer. 29 The conjugation or encapsulation of these low<br />

molecular weight drugs within polymeric nanoparticles has distinct advantages: Increased plasma<br />

half-life of low molecular weight drugs; increased solubility of hydrophobic drugs; and controlled<br />

release of the drug at lower doses. The targeting of these nanoparticles to the tumor tissues is<br />

achieved either passively <strong>by</strong> the EPR effect or actively <strong>by</strong> surface conjugating a tumor-specific<br />

ligand. 30<br />

13.4 NANOPARTICLES FOR GENE DELIVERY<br />

Blood Capillary<br />

Endothelial Cells<br />

Leaky Vasculature<br />

FIGURE 13.1 Schematic of enhanced permeability and retention effect. Passive targeting to the tumor cells is<br />

achieved <strong>by</strong> the extravasation of the polymeric nanoparticles through the leaky vasculature of the blood<br />

capillaries in the tumor. The leaky vasculature, along with the lack of lymphatic drainage, is called the<br />

EPR effect and is used to passively target nanocarriers to the tumor.<br />

Due to the complexity and heterogeneity in cancer, gene therapy can provide a unique approach for<br />

treatment. Research on gene therapy focuses on the successful in vivo transfer of genes to the target<br />

tissue for the sustained expression of the genes of dysfunction. Various vectors, both viral and nonviral<br />

in origin, are used for this purpose. 31–34 The limitations associated with viral vectors, namely<br />

issues of integration with the host genome, self-replication, recombination potential, and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


234<br />

immunogenicity, have encouraged researchers to seek alternatives such as nanoparticles, liposomes,<br />

and other cationic complexes. 35–37 Despite the low transfection efficiency associated with<br />

these non-viral vectors, they have the advantages of safety and high encapsulation efficiency. The<br />

DNA complexes of cationic liposomes and polymers have high transfection efficiency in vitro;<br />

however, owing to their toxicity and rapid clearance, these formulations have limited efficiency<br />

in vivo. The cationic liposomes and polymers, being positively charged, form aggregates with the<br />

negatively charged serum proteins and opsonins, resulting in enhanced phagocytosis and clearance<br />

from the blood circulation. 38<br />

In addition to cationic polymers, several other natural and synthetic polymers have been used in<br />

preparing DNA-encapsulated nanoparticles. Polymeric nanoparticles are increasingly popular,<br />

owing to such advantages as their small size, ease of production, and administration. Several<br />

polymers have been investigated as vectors for gene delivery applications. There has been a fair<br />

amount of success in reducing immunogenicity and cytotoxicity with the concomitant enhancement<br />

of the efficiency of transfection with these polymers. Besides protecting the DNA from degradation,<br />

the nanoparticles enhance the targeting of tumor tissues. 38<br />

13.5 LONG-CIRCULATING NANOPARTICLES<br />

Long-circulating nanoparticles can be created through surface modification of conventional nanoparticles<br />

with water-soluble polymers such as polyethylene glycol (PEG) or polyethylene oxide<br />

(PEO) (Figure 13.2). The hydrophilic nature of these surface modifiers minimizes the interactions<br />

between the nanoparticles and plasma proteins (opsonins), resulting in reduced uptake <strong>by</strong> the<br />

reticulo-endothelial system (RES). The major outcome of modification with PEG or other hydrophilic<br />

flexible polymers is a significant increase in circulation time, the advantages of which include<br />

maintenance of optimal therapeutic concentration of the drug in the blood after a single administration<br />

of the drug carrier, increased probability of extravasation and retention of the colloidal<br />

carrier in areas of discontinuous endothelium, and enhancement in targetability of the system <strong>by</strong> use<br />

of a target-specific ligand. 39<br />

The protective action (Stealth w property) of PEG is mainly due to the formation of a dense,<br />

hydrophilic cloud of long polyethylene chains on the surface of the colloidal particle that reduces<br />

the hydrophobic interactions with the RES. The tethered or chemically anchored PEG chains can<br />

(a)<br />

PEG PPO<br />

(b)<br />

Nanotechnology for Cancer Therapy<br />

PEO PEO<br />

FIGURE 13.2 Schematic of long-circulating polymeric nanoparticles. The polymeric nanoparticles are made<br />

long-circulating <strong>by</strong> surface modification. The nanoparticles prepared from a hydrophilic polymer are modified<br />

using poly(ethylene glycol) (a), and the nanoparticles prepared from hydrophobic polymers are modified using<br />

Pluronic w , a triblock co-polymer of poly(ethylene oxide) and poly(propylene oxide) (b).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Long-Circulating Polymeric Nanoparticles for Drug and Gene Delivery to Tumors 235<br />

undergo spatial conformations, there<strong>by</strong> preventing the opsonization of particles <strong>by</strong> the RES of the<br />

liver and spleen and improving the circulation time of molecules and particles in the blood. The<br />

greater the flexibility of the polymer, the greater the total number of possible conformations and<br />

transitions from one conformation to another. 40–43 Water molecules form a structured shell through<br />

hydrogen bonding to the ether oxygens of PEG. The tightly bound water around PEG chains forms a<br />

hydrated film around the particle and prevents protein interactions. 44 Furthermore, PEGylation may<br />

also increase the hydrodynamic size of the particles, decreasing their clearance through the kidneys,<br />

renal filtration being dependent on molecular mass and volume. This would ultimately result in an<br />

increase in the circulation half-life of the particles. 45,46<br />

The size, molecular weight, and shape of the PEG fraction and the linkage used to connect it to<br />

the entity of interest determine the consequences of PEGylation in relation to protein adsorption<br />

and pharmacokinetics such as volume of distribution, circulation time, and renal clearance. When<br />

formulated into colloidal particles, the PEG density on the colloidal surface can be changed <strong>by</strong><br />

using PEG of appropriate molecular weight (PEG chain length) and molar ratio (the grafting<br />

efficiency). Longer PEG chains offer greater steric influence around the colloidal entity, similar<br />

to increased grafting density with shorter PEG chains. Longer PEG chains may also collapse onto<br />

the nanoparticle surface, providing a hydrophilic shield. 40<br />

Besides PEG, other hydrophilic polymers, including polyvinyl alcohol, polyacryl amide, polyvinyl<br />

pyrrolidone, poly-[N-(2-hydroxypropyl)methacrylamide], polysorbate-80, and block<br />

co-polymers such as poloxomer (Pluronic w ) and poloxamine (Tetronic w ), are also being used to<br />

modify the physicochemical properties of the colloidal carriers. 32,47,48<br />

The polymeric nanoparticles modified with PEG or PEO are mainly used to passively target<br />

tumors through the EPR effect. The surface-modified long-circulating polymeric nanoparticles are<br />

used to deliver both genes and drugs to the tumor tissues. Some of the polymeric nanoparticulate<br />

systems that were developed <strong>by</strong> this group will be discussed in the sections that follow.<br />

13.6 ILLUSTRATIVE EXAMPLES OF POLYMERIC NANOPARTICLES<br />

FOR DRUG DELIVERY<br />

13.6.1 POLYETHYLENE OXIDE-MODIFIED POLY(b-AMINO ESTER) NANOPARTICLES<br />

A representative biodegradable, hydrophobic poly(b-amino ester) (PBAE) with pHsensitive<br />

solubility properties was synthesized <strong>by</strong> conjugate addition of 4,4 0 -trimethylenedipiperidine with<br />

1,4-butanediol diacrylate, developed in Professor Robert Langer’s lab at Massachusetts Institute of<br />

Technology. The paclitaxel-loaded nanoparticles (150–200 nm) prepared from PBAE were<br />

modified with Pluronic w F-108 (poloxamer 407), a triblock copolymer of polyethylene oxide/polypropylene<br />

oxide/polyethylene oxide (PEO/PPO/PEO). The PPO segment of the triblock polymer<br />

attaches to the hydrophobic surface of the nanoparticles, and the hydrophilic PEO segment contributes<br />

to the stealth properties of the polymeric nanoparticles. The pH-sensitive nature of the particles<br />

prepared from PBAE has already been shown <strong>by</strong> in vitro release studies carried out in the presence<br />

of buffers of pH ranging from 5.0 to 7.4 and was found to rapidly degrade in a medium of pH less<br />

than 6.5. Therefore, these nanoparticles were expected to readily release their contents within the<br />

acidic tumor microenvironment and in the endosomes and lysosomes of the cells upon internalization.<br />

This was confirmed <strong>by</strong> the in vitro cellular uptake of the PEO–PBAE nanoparticles<br />

encapsulated with tritiated [ 3 H]-paclitaxel <strong>by</strong> human breast adenocarcinoma cells (MDA-<br />

MB231). 49,50<br />

The biodistribution of these PEO-modified PBAE nanoparticles was carried out <strong>by</strong> encapsulating<br />

a lipophilic form of the radionuclide indium-111 ( 111 indium oxine). Following tail vein<br />

injection in nude mice bearing a human ovarian xenograft, the radiolabeled PEO-modified<br />

PBAE nanoparticles were found to accumulate in the highly perfused organs such as the liver,<br />

spleen, and lungs with greater entrapment in the microvasculature of the lungs during the initial<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


236<br />

Nanotechnology for Cancer Therapy<br />

time points. The increasing concentrations in the kidney also indicate that the nanoparticles, once<br />

internalized, were disintegrated and eliminated through the kidney. The plasma half-life of the<br />

unmodified nanoparticles was reported to be one to ten minutes. By virtue of surface modification<br />

with PEO, the PBAE nanoparticles were shown to have improved circulation times, resulting in a<br />

mean residence time in the systemic circulation of 21 h. The paclitaxel-encapsulated PEO–PBAE<br />

nanocarriers were found to deliver the drug efficiently to solid tumors, resulting in a 5.2-fold and<br />

23-fold higher concentration of the drug at one hour and five hours post-administration relative to<br />

the solution form of the drug. 40 From the tumor accumulation of the paclitaxel-loaded ( 3 H-labeled)<br />

nanoparticles, it is evident that the pH-sensitive PEO–PBAE nanoparticle formulations can deliver<br />

significantly higher concentrations of the drug into the tumor than the solution form.<br />

13.6.2 POLY(ETHYLENE OXIDE)-MODIFIED POLY(3-CAPROLACTONE) NANOPARTICLES<br />

Poly(3-caprolactone) (PCL) is another biodegradable polymer that has been used to encapsulate<br />

hydrophobic drugs. Using this polymer, nanoparticles were prepared <strong>by</strong> solvent displacement in an<br />

acetone-water system in the presence of Pluronic w . The solvent displacement technique used for<br />

the preparation of PCL nanoparticles facilitates instant adsorption of PPO–PEO groups when the<br />

organic solution of the polymer is introduced into aqueous solution containing the stabilizer. 51 In<br />

addition, it also favors the encapsulation of hydrophobic drugs such as tamoxifen that could be<br />

dissolved along with the polymer in the organic phase, resulting in a high entrapment efficiency of<br />

greater than 90% at loading levels of 20% of the weight of the drug. The intracellular uptake of<br />

these nanoparticles in MCF-7 estrogen receptor-positive breast cancer cells and MDA-MB231<br />

human breast adenocarcinoma cells was monitored at different time points using tritiated [ 3 H]tamoxifen.<br />

The results showed that the cell uptake followed saturable kinetics with most of the<br />

nanoparticles being internalized within the first 30 min of incubation. 52<br />

The in vivo disposition of these PEO-modified PCL nanoparticles was completed in mice<br />

bearing MDA-MB231 xenograft breast cancer tumors as it is a well-characterized and simpler<br />

model compared to MCF-7 that requires estrogen priming for growth. This study was used to<br />

compare the biodistribution profiles of the nanoparticles modified <strong>by</strong> the Pluronic w F-68 containing<br />

30 residues of propylene oxide (PO) and 76 residues of ethylene oxide (EO) with those of Pluronic w<br />

F-108 that has 56 PO residues and 122 EO residues. The nanoparticulate formulations encapsulated<br />

with radiolabled tamoxifen were used. As expected, upon intravenous administration, the Pluronic w -<br />

modified nanoparticles had higher tumor concentrations in comparision to the tamoxifen solution<br />

and drug-loaded, unmodified nanoparticles. At early time points (one hour), the nanoparticles<br />

modified with Pluronic w F-108 had greater concentration in the tumor with no significant difference<br />

in the concentration of the Pluronic w -modified formulations at six hours post-injection. A similar<br />

trend has been observed in plasma concentration-time profiles with PEO–PCL formulations circulating<br />

for a longer time than the controls. 40<br />

In a parallel study, the PEO-modified PCL nanoparticles were radiolabeled <strong>by</strong> a similar<br />

procedure specific to PEO–PBAE nanoparticles. The nanoparticles, encapsulated with [ 3 H]<br />

tritium-labeled paclitaxel, were used to understand the change in concentration and localization<br />

of the drug in ovarian tumors (SKOV3). From the biodistribution studies, it was shown that the<br />

modification of PCL nanoparticles with PEO had extended the mean residence time to up to 25 h.<br />

Hydrophobic drugs such as paclitaxel were found to have high plasma concentrations as a result of<br />

their protein binding capacity; however, they were cleared from the blood within 24 h. The<br />

circulation time of such drugs has been enhanced <strong>by</strong> encapsulating them in PEO–PCL nanoparticles<br />

that, in turn, has resulted in higher concentrations of the drug in the tumors. The PEO–PCL<br />

nanoparticles have resulted in an 8.7-fold increase in drug concentration at five-hour time points<br />

when compared to the solution form of the drug. 40<br />

In another study, PEO-modified PCL nanoparticles were used for combination therapy of<br />

ceramide with paclitaxel in order to overcome MDR, particularly in cases of breast and ovarian<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Long-Circulating Polymeric Nanoparticles for Drug and Gene Delivery to Tumors 237<br />

cancers. Several MDR specimens of cancer were found to exhibit elevated levels of the enzyme<br />

glucosylceramide synthease (GCS) (also called ceramide glucosyl transferase) that is responsible<br />

for the inactivation of ceramide, a messenger in apoptotic signaling to its non-functional moiety<br />

glucosylceramide. 53–55 These findings suggest the importance of the role of ceramide in the<br />

mediation of a cytotoxic response and its function as an apoptotic messenger in the signaling<br />

pathway.<br />

To potentially overcome MDR in ovarian cancer cell lines, C6-ceramide has been encapsulated<br />

along with paclitaxel into PEO-modified PCL nanoparticles. Upon treatment of the cells with<br />

paclitaxel, the MDR cell line SKOV3/TR exhibited 65.65G2.16% viability at 1 mM dose; the<br />

sensitive cell line SKOV3 showed 16.37G0.41% viability at 100 nM dose. Co-treatment of<br />

these cells along with 20 mM C6-ceramide in addition to paclitaxel (1 mM in the case of a resistant<br />

cell line and 100 nM in a sensitive cell line) resulted in a cell viability of 2.69G0.51% with the<br />

resistant cells and 7.38G1.25% with the sensitive cell lines, indicating a significant increase in cell<br />

death when compared to the paclitaxel treatment alone. Furthermore, the co-encapsulation of these<br />

drugs within PEO–PCL nanoparticles resulted in enhanced cell kill compared to the drugs alone.<br />

A 10 nM dose of paclitaxel, delivered in combination with ceramide in PEO–PCL nanoparticles,<br />

resulted in 63.98G4.9% viability, and the free drugs in solution at these doses did not provoke any<br />

cell kill in the resistant cell line. The use of these drug-loaded nanoparticles resulted in a 100-fold<br />

increase in chemosensitivity of the MDR cells. These results demonstrate the clinical use of<br />

PEO–PCL nanoparticles in overcoming MDR <strong>by</strong> combination therapy.<br />

13.7 ILLUSTRATIVE EXAMPLES OF POLYMERIC NANOPARTICLES<br />

FOR GENE DELIVERY<br />

13.7.1 POLYETHYLENE GLYCOL-MODIFIED GELATIN NANOPARTICLES<br />

To develop safe and effective systemically administered non-viral gene therapy vectors for solid<br />

tumors, PEGylated gelatin was synthesized and fabricated into nanoparticles. The PEG-modified<br />

gelatin was synthesized <strong>by</strong> reacting Type-B gelatin with PEG-epoxide that was further confirmed<br />

<strong>by</strong> electron spectroscopy for chemical analysis (ESCA), the results of which had shown an increase<br />

in the presence of an ether carbon (–C–O–) peak in the PEGylated gelatin nanoparticles. The<br />

experiments performed using these polymeric nanoparticulate carriers proved their ability to encapsulate<br />

DNA, improved transfection efficiency in vitro, and enhanced the biodistribution profile<br />

when the carriers were injected into mice bearing tumors. 56–58<br />

The in vitro cytotoxicity assays indicate that both gelatin and PEGylated gelatin are non-toxic<br />

to the cells. The cell uptake and trafficking studies of the control (gelatin) and PEGylated gelatin<br />

nanoparticles encapsulated with electron-dense gold nanoparticles confirmed that the particles were<br />

internalized <strong>by</strong> an endocytic pathway and remained stable during the vesicular transport. Further,<br />

the transfection studies carried out using the DNA (pEGFP-N1) encapsulated nanoparticles were<br />

internalized in NIH-3T3 fibroblast cells within the first six hours of incubation. Green fluorescent<br />

protein expression was observed after 12 h of nanoparticle incubation and remained stable for up to<br />

96 hours. Flow cytometry results showed that the DNA transfection efficiency of PEGylated gelatin<br />

nanoparticles was better than that of gelatin. 57<br />

In order to determine the biodistribution profile, nanoparticles were radiolabeled with iodine-<br />

125 [ 125 I] and intravenously injected into mice bearing Lewis lung carcinoma (<strong>LLC</strong>) tumors. From<br />

the radioactivity of the plasma and other organs collected, it was evident that the majority of the<br />

PEGylated gelatin nanoparticles remained in the blood pool or were taken up <strong>by</strong> the tumor mass and<br />

liver. A two-fold increase in concentration of the PEGylated nanoparticles was observed in plasma<br />

even at three-hour time points, and about 4–5% of the recovered dose of PEGylated gelatin<br />

nanoparticles was found to be present in the tumor mass for up to 12 h. Upon non-compartmental<br />

pharmacokinetic analysis of the plasma and tumor concentration-time profiles, it was observed<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


238<br />

that PEGylated gelatin nanoparticles had greater mean residence time in plasma and a more than<br />

six-fold increase in half-life in the tumor. The results of this study showed the stealth nature of<br />

the PEGylated gelatin nanoparticles in avoiding uptake <strong>by</strong> RES, allowing the nanoparticles to<br />

circulate longer in plasma. 59<br />

TheinvivotransfectionefficiencyofplasmidDNA-encapsulated(pCMV-b encoding<br />

b-galactosidase) PEGylated gelatin nanoparticles was evaluated <strong>by</strong> injecting these nanoparticles<br />

intravenously into <strong>LLC</strong> tumor-bearing female C57BL/6J mice. Following systemic administration,<br />

the animals were sacrificed, and the transgene expression in different organs was quantitatively<br />

determined <strong>by</strong> using o-nitrophenyl-b-d-galactopyranosidase (ONPG), a clear substrate of<br />

b-galactosidase that is converted to a yellow-colored product with an absorbance maximum at<br />

420 nm, and qualitatively <strong>by</strong> X-gal w tissue staining. When compared to the gelatin nanoparticles,<br />

the PEGylated gelatin nanoparticles were found to efficiently transfect the tumor cells with the<br />

b-galactosidase expression increasing at time points up to 96 h post-transfection with the absorbance<br />

value at 420 nm increasing from 0.6 to 0.85 starting from 12 h post-transfection. The results<br />

of these studies clearly indicate that the long circulating, biocompatible, and biodegradable nanoparticles<br />

of PEGylated gelatin would be ideal for gene delivery applications to solid tumors. 60<br />

13.7.2 POLYETHYLENE GLYCOL-MODIFIED THIOLATED-GELATIN NANOPARTICLES<br />

Another study synthesized thiolated gelatin and formulated it into nanoparticles that could rapidly<br />

release their payload in the presence of a high concentration of glutathione. The thiolated gelatins<br />

with different degrees of thiolation were characterized <strong>by</strong> Ellman’s assay to determine degree of<br />

thiolation and <strong>by</strong> in vitro cytotoxicity assay. Upon carrying out the qualitative and quantitative<br />

transfection studies, the nanoparticles of thiolated gelatin, made with 20 mg of aminothiolane per<br />

gram of gelatin (SHGel-20) and encapsulated with plasmid DNA (pEGFP-N1), were found to have<br />

higher transfection efficiency in mouse fibroblast (NIH-3T3) cells in vitro (Figure 13.3). 61<br />

These nanoparticles, prepared <strong>by</strong> the desolvation method, were post-PEGylated with polyethylene<br />

glycol-succinimidyl glutarate (PEG-SG) of molecular weight 2,000 Da. The radiolabeled<br />

( 111 Indium) long-circulating (PEG-modified) thiolated-gelatin nanoparticles were injected intravenously<br />

into mice bearing human breast cancer (MDA-MB435) xenografts. The biodistribution and<br />

pharmacokinetics of these particles were compared to those of nanoparticles prepared from thiolated-gelatin,<br />

gelatin, and PEG-modified gelatin. The unmodified gelatin and thiolated gelatin<br />

nanoparticles were shown to have a plasma half-life of approximately three hours. Upon modification<br />

with PEG, the nanoparticles were found to have prolonged circulation times extending up to<br />

10.7 h in the case of PEG-gelatin and 15.3 h with PEG-thiolated-gelatin nanoparticles. The<br />

successful surface hydrophilization of the gelatin and thiolated gelatin nanoparticles resulted in<br />

enhanced circulation times with almost 7% of the dose remaining in plasma at 12 h. The results also<br />

showed that the thiolated-gelatin nanoparticles (both non-modified and PEG-modified) had greater<br />

accumulations in the tumor than the nanoparticles prepared from non-thiolated gelatin. These<br />

thiolated-gelatin nanoparticles that are an improvement over gelatin nanoparticles can be made<br />

long-circulating <strong>by</strong> surface modification with PEG, and they can be used for rapid delivery of the<br />

payload the tumor tissues. The mechanisms of long-circulation through steric hindrance when in<br />

circulation and preferential accumulation in the tumor mass or inside the cells as a result of the cleavage<br />

of the disulfide bonds (arising from thiolation) in the highly reductive tumor-microenvironment<br />

can be exploited for the delivery of drugs and genes in clinical applications. 40,61<br />

13.8 CONCLUDING REMARKS<br />

Nanotechnology for Cancer Therapy<br />

There are multiple factors affecting the delivery of drugs and genes to tumors. Factors such as blood<br />

flow, angiogenesis, microvessel density, interstitial pressure, macrophage activity, extracellular and<br />

intracellular components, and, most importantly, the physicochemical properties of the drug carrier<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Long-Circulating Polymeric Nanoparticles for Drug and Gene Delivery to Tumors 239<br />

play an important role in the transport of drugs and macromolecules to tumors. This lab is particularly<br />

interested in enhancing the transport of drugs and genes to tumors <strong>by</strong> altering the<br />

physicochemical properties of the polymeric carrier used to encapsulate the drugs and genes. As<br />

shown in the illustrative examples, the biodistribution properties of the polymeric carriers could be<br />

modified using hydrophilic polymers such as PEG and PEO. This therapeutic strategy could be used<br />

to alter the passive/active targeting ability of the drug and gene carriers. However, the delivery of<br />

these newer agents is still a challenge, highlighting the necessity of additional research in this area.<br />

ACKNOWLEDGMENTS<br />

These studies have been supported <strong>by</strong> grants R01-CA095522 and R01-CA119617 from the<br />

National Cancer Institute, National Institutes of Health. Our research work on poly(beta-amino<br />

ester) nanoparticles is done in collaboration with Professor Robert Langer’s lab at MIT<br />

(Cambridge, MA). We deeply appreciate the on-going, highly productive collaborations with<br />

his group and especially the assistance of Dr. Steven Little, Dr. Daniel Anderson, and<br />

Mr. David Nguyen.<br />

REFERENCES<br />

(a)<br />

FIGURE 13.3 Transfection images of mouse fibroblast (NIH3T3) cells treated with thiolated-gelatin nanoparticles<br />

encapsulated with plasmid DNA (pEGFP-N1). The images show fluorescence of the expressed green<br />

fluorescent protein at 6 h (a) and 96 h post-transfection (b).<br />

(b)<br />

1. Boehm, T., Folkman, J., Browder, T., and O’Reilly, M. S., Antiangiogenic therapy of experimental<br />

cancer does not induce acquired drug resistance, Nature, 390(6658), 404–407, 1997.<br />

2. Kerbel, R. S., A cancer therapy resistant to resistance, Nature, 390(6658), 335–336, 1997.<br />

3. Kirsch, M., Strasser, J., Allende, R., Bello, L., Zhang, J., and Black, P. M., Angiostatin suppresses<br />

malignant glioma growth in vivo, Cancer Research, 58(20), 4654–4659, 1998.<br />

4. Ganjavi, H., Gee, M., Narendran, A., Freedman, M. H., and Malkin, D., Adenovirus-mediated p53<br />

gene therapy in pediatric soft-tissue sarcoma cell lines: Sensitization to cisplatin and doxorubicin,<br />

Cancer Gene Therapy, 12(4), 397–406, 2005.<br />

5. Liu, Y., Huang, H., Saxena, A., and Xiang, J., Intratumoral coinjection of two adenoviral vectors<br />

expressing functional interleukin-18 and inducible protein-10, respectively, synergizes to facilitate<br />

regression of established tumors, Cancer Gene Therapy, 9(6), 533–542, 2002.<br />

6. Van Meir, E. G., Polverini, P. J., Chazin, V. R., Su Huang, H. J., de Tribolet, N., and Cavenee, W. K.,<br />

Release of an inhibitor of angiogenesis upon induction of wild type p53 expression in glioblastoma<br />

cells, Nature Genetics, 8(2), 171–176, 1994.<br />

7. Schiffelers, R. M., Ansari, A., Xu, J., Zhou, Q., Tang, Q., Storm, G., Molema, G., Lu, P. Y., Scaria, P.<br />

V., Woodle, M. C. Cancer siRNA therapy <strong>by</strong> tumor selective delivery with ligand-targeted sterically<br />

stabilized nanoparticle Nucleic Acids Research, 32(19), e149, 2004 (http://nar.oxfordjournal.org/cgi/<br />

content/abstract/32/19/e149).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


240<br />

Nanotechnology for Cancer Therapy<br />

8. Yuan, F., Salehi, H. A., Boucher, Y., Vasthare, U. S., Tuma, R. F., and Jain, R. K., Vascular<br />

permeability and microcirculation of gliomas and mammary carcinomas transplanted in rat and<br />

mouse cranial windows, Cancer Research, 54(17), 4564–4568, 1994.<br />

9. Yuan, F., Leunig, M., Huang, S. K., Berk, D. A., Papahadjopoulos, D., and Jain, R. K., Microvascular<br />

permeability and interstitial penetration of sterically stabilized (stealth) liposomes in a human tumor<br />

xenograft, Cancer Research, 54(13), 3352–3356, 1994.<br />

10. Jain, R. K., Physiological barriers to delivery of monoclonal antibodies and other macromolecules in<br />

tumors, Cancer Research, 50(Suppl. 3), 814s–819s, 1990.<br />

11. Jang, S. H., Wientjes, M. G., Lu, D., and Au, J. L., Drug delivery and transport to solid tumors,<br />

Pharmaceutical Research, 20(9), 1337–1350, 2003.<br />

12. Maeda, H., SMANCS and polymer-conjugated macromolecular drugs: Advantages in cancer<br />

chemotherapy, Advanced Drug Delivery Reviews, 46(1–3), 169–185, 2001.<br />

13. Maeda, H. and Matsumura, Y., Tumoritropic and lymphotropic principles of macromolecular drugs,<br />

Critical Reviews in Therapeutic Drug Carrier Systems, 6(3), 193–210, 1989.<br />

14. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the anti-tumor agent<br />

smancs, Cancer Research, 46(12 Pt 1), 6387–6392, 1986.<br />

15. Balkwill, F. and Mantovani, A., Inflammation and cancer: Back to Virchow?, Lancet, 357(9255),<br />

539–545, 2001.<br />

16. Coussens, L. M. and Werb, Z., Inflammation and cancer, Nature, 420(6917), 860–867, 2002.<br />

17. Lewis, C. and Murdoch, C., Macrophage responses to hypoxia: Implications for tumor progression and<br />

anti-cancer therapies, American Journal of Pathology, 167(3), 627–635, 2005.<br />

18. Mantovani, A., Sozzani, S., Locati, M., Allavena, P., and Sica, A., Macrophage polarization: Tumorassociated<br />

macrophages as a paradigm for polarized M2 mononuclear phagocytes, Trends in<br />

Immunology, 23(11), 549–555, 2002.<br />

19. Luo, D., Haverstick, K., Belcheva, N., Han, E., and Saltzman, M., Poly(ethylene glycol)-conjugated<br />

PAMAM dendrimer for biocompatible, high-efficiency DNA delivery, Macromolecules, 35,<br />

3456–3462, 2002.<br />

20. Kircheis, R., Blessing, T., Brunner, S., Wightman, L., and Wagner, E., Tumor targeting with surfaceshielded<br />

ligand—polycation DNA complexes, Journal of Controlled Release, 72(1–3), 165–170,<br />

2001.<br />

21. Kim, S. H., Jeong, J. H., Cho, K. C., Kim, S. W., and Park, T. G., Target-specific gene silencing <strong>by</strong><br />

siRNA plasmid DNA complexed with folate-modified poly(ethylenimine), Journal of Controlled<br />

Release, 104, 223–232, 2005.<br />

22. Suh, W., Han, S. O., Yu, L., and Kim, S. W., An angiogenic, endothelial-cell-targeted polymeric gene<br />

carrier, Molecular Therapy, 6(5), 664–672, 2002.<br />

23. Kursa, M., Walker, G. F., Roessler, V., Ogris, M., Roedl, W., Kircheis, R., and Wagner, E., Novel<br />

shielded transferrin-polyethylene glycol-polyethylenimine/DNA complexes for systemic tumortargeted<br />

gene transfer, Bioconjugate Chemistry, 14(1), 222–231, 2003.<br />

24. Lee, H., Kim, T. H., and Park, T. G., A receptor-mediated gene delivery system using streptavidin and<br />

biotin-derivatized, pegylated epidermal growth factor, Journal of Controlled Release, 83(1), 109–119,<br />

2002.<br />

25. Merdan, T., Callahan, J., Petersen, H., Kunath, K., Bakowsky, U., Kopeckova, P., Kissel, T., and<br />

Kopecek, J., Pegylated polyethylenimine-Fab’ antibody fragment conjugates for targeted gene<br />

delivery to human ovarian carcinoma cells, Bioconjugate Chemistry, 14(5), 989–996, 2003.<br />

26. Natarajan, A., Xiong, C. Y., Albrecht, H., DeNardo, G. L., and DeNardo, S. J., Characterization of<br />

site-specific ScFv PEGylation for tumor-targeting pharmaceuticals, Bioconjugate Chemistry, 16(1),<br />

113–121, 2005.<br />

27. Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancer therapy,<br />

Advanced Drug Delivery Reviews, 56(11), 1649–1659, 2004.<br />

28. Ehrlich, P., Studies in Immunity, Wiley, New York, 1906.<br />

29. Ringsdorf, H., Structure and properties of pharmacologically active polymers, Journal of Polymer<br />

Science Polymer Symposium, 51, 135–153, 1975.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Long-Circulating Polymeric Nanoparticles for Drug and Gene Delivery to Tumors 241<br />

30. Stastny, M., Strohalm, J., Plocova, D., Ulbrich, K., and Rihova, B., A possibility to overcome<br />

p-glycoprotein (PGP)-mediated multidrug resistance <strong>by</strong> antibody-targeted drugs conjugated to<br />

N-(2-hydroxypropyl)methacrylamide (HPMA) copolymer carrier, European Journal of Cancer, 35,<br />

459–466, 1999.<br />

31. El-Aneed, A., An overview of current delivery systems in cancer gene therapy, Journal of Controlled<br />

Release, 94(1), 1–14, 2004.<br />

32. Fenske, D. B., MacLachlan, I., and Cullis, P. R., Long-circulating vectors for the systemic delivery of<br />

genes, Current Opinion in Molecular Therapeutics, 3(2), 153–158, 2001.<br />

33. Han, S., Mahato, R. I., Sung, Y. K., and Kim, S. W., Development of biomaterials for gene therapy,<br />

Moleular Therapeutics, 2(4), 302–317, 2000.<br />

34. McCormick, F., Cancer gene therapy: Fringe or cutting edge?, Nature Reviews Cancer, 1(2),<br />

130–141, 2001.<br />

35. Lehrman, S., Virus treatment questioned after gene therapy death, Nature, 401(6753), 517–518, 1999.<br />

36. Marshall, E., Gene therapy. Second child in French trial is found to have leukemia, Science,<br />

299(5605), 320, 2003.<br />

37. Thomas, C. E., Ehrhardt, A., and Kay, M. A., Progress and problems with the use of viral vectors for<br />

gene therapy, Nature Reviews Genetics, 4(5), 346–358, 2003.<br />

38. Kommareddy, S., Tiwari, S. B., and Amiji, M. M., Long-circulating polymeric nanovectors for<br />

tumor-selective gene delivery, Technology in Cancer Research and Treatment, 4(6), 615–626, 2005.<br />

39. Torchilin, V. P., Drug targeting, European Journal of Pharmaceutial Sciences, 11(Suppl. 2),<br />

S81–S91, 2000.<br />

40. Kommareddy, S., Shenoy, D. B., and Amiji, M., Nanoparticulate carriers of gelatin and gelatin<br />

derivatives, In Biological and Pharmaceutical Nanomaterials, Kumar, C., Ed., Wiley-VCH, Weinheim,<br />

Germany, pp. 330–347, 2005.<br />

41. Torchilin, V. P., Polymer-coated long-circulating microparticulate pharmaceuticals, Journal of<br />

Microencapsulation, 15(1), 1–19, 1998.<br />

42. Torchilin, V. P. and Papisov, M. I., Why do polyethylene glycol-coated liposomes circulate so long?,<br />

Journal of Liposome Research, 4(1), 725–739, 1994.<br />

43. Zalipsky, S., Functionalized Poly(ethylene glycol) for preparation of biologically revalant conjugates,<br />

Bioconjugate Chemistry, 6, 150–165, 1995.<br />

44. Gref, R., Domb, A. J., Quellec, P., Blunk, T., Müller, R. H., Verbavatz, J. M., and Langer, R., The<br />

controlled intravenous delivery of drugs using PEG-coated sterically stabilized nanospheres,<br />

Advanced Drug Delivery Reviews, 16(2–3), 215–233, 1995.<br />

45. Delgado, C., <strong>Francis</strong>, G. E., and Fisher, D., The uses and properties of PEG-linked proteins, Critical<br />

Reviews in Therapeutic Drug Carrier Systems, 9(3–4), 249–304, 1992.<br />

46. Mehvar, R., Dextrans for targeted and sustained delivery of therapeutic and imaging agents, Journal<br />

of Controlled Release, 69(1), 1–25, 2000.<br />

47. Torchilin, V. P., How do polymers prolong circulation time of liposomes?, Journal of Liposome<br />

Research, 6(1), 99–116, 1996.<br />

48. Oupicky, D., Howard, K. A., Konak, C., Dash, P. R., Ulbrich, K., and Seymour, L. W., Steric<br />

stabilization of poly-L-Lysine/DNA complexes <strong>by</strong> the covalent attachment of semitelechelic<br />

poly[N-(2-hydroxypropyl)methacrylamide], Bioconjugate Chemistry, 11(4), 492–501, 2000.<br />

49. Lynn, D. M., Amiji, M. M., and Langer, R., pH-responsive polymer microspheres: Rapid release of<br />

encapsulated material within the range of intracellular pH, Angewandte Chemie (International ed. in<br />

English), 40(9), 1707–1710, 2001.<br />

50. Potineni, A., Lynn, D. M., Langer, R., and Amiji, M. M., Poly(ethylene oxide)-modified<br />

poly(beta-amino ester) nanoparticles as a pH-sensitive biodegradable system for paclitaxel delivery,<br />

Journal of Controlled Release, 86(2–3), 223–234, 2003.<br />

51. Scholes, P. D., Coombes, A. G., Illum, L., Davis, S. S., Watts, J. F., Ustariz, C., Vert, M., and Davies,<br />

M. C., Detection and determination of surface levels of poloxamer and PVA surfactant on biodegradable<br />

nanospheres using SSIMS and XPS, Journal of Controlled Release, 59(3), 261–278, 1999.<br />

52. Chawla, J. S. and Amiji, M. M., Cellular uptake and concentrations of tamoxifen upon administration<br />

in poly(epsilon-caprolactone) nanoparticles, AAPS PharmSci, 5(1), E3, 2003 (http://www.aapsj.org/<br />

view.asp?art=ps050103).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


242<br />

Nanotechnology for Cancer Therapy<br />

53. Lavie, Y., Cao, H., Bursten, S. L., Giuliano, A. E., and Cabot, M. C., Accumulation of glucosylceramides<br />

in multidrug-resistant cancer cells, Journal of Biological Chemistry, 271, 19530–19536, 1996.<br />

54. Liu, Y. Y., Han, T. Y., Giuliano, A. E., and Cabot, M. C., Ceramide glycosylation potentiates cellular<br />

multidrug resistance, FASEB Journal, 15(3), 719–730, 2001.<br />

55. Morjani, H., Aouali, N., Belhoussine, R., Veldman, R. J., Levade, T., and Manfait, M., Elevation of<br />

glucosylceramide in multidrug-resistant cancer cells and accumulation in cytoplasmic droplets,<br />

International Journal of Cancer, 94, 157–165, 2001.<br />

56. Kaul, G. and Amiji, M., Long-circulating poly(ethylene glycol)-modified gelatin nanoparticles for<br />

intracellular delivery, Pharmaceutical Research, 19(7), 1062–1068, 2002.<br />

57. Kaul, G. and Amiji, M., Cellular interactions and in vitro DNA transfection studies with poly(ethylene<br />

glycol)-modified gelatin nanoparticles, Journal of Pharmaceutical Sciences, 94(1), 184–198, 2004.<br />

58. Kaul, G., Lee-Parsons, C., and Amiji, M., Poly(ethylene glycol)-modified gelatin nanoparticles for<br />

intracellular delivery, Pharmaceutical Engineers, 23(5), 1–5, 2003.<br />

59. Kaul, G. and Amiji, M., Biodistribution and targeting potential of poly(ethylene glycol)-modified<br />

gelatin nanoparticles in subcutaneous murine tumor model, Journal of Drug Targeting, 12(9–10),<br />

585–591, 2004.<br />

60. Kaul, G. and Amiji, M., Tumor-targeted gene delivery using poly(ethylene glycol)-modified gelatin<br />

nanoparticles: In vitro and in vivo studies, Pharmaceutical Research, 22(6), 951–961, 2005.<br />

61. Kommareddy, S. and Amiji, M., Preparation and evaluation of thiol-modified gelatin nanoparticles for<br />

intracellular DNA delivery in response to glutathione, Bioconjugate Chemistry, 16(6), 1423–1432,<br />

2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


14<br />

CONTENTS<br />

Biodegradable PLGA/PLA<br />

Nanoparticles for Anti-Cancer<br />

Therapy<br />

Sanjeeb K. Sahoo and Vinod Labhasetwar<br />

14.1 Introduction ....................................................................................................................... 243<br />

14.2 General Principles of Drug Targeting to Cancer.............................................................. 244<br />

14.2.1 Passive Targeting ................................................................................................ 244<br />

14.2.2 Active Targeting.................................................................................................. 244<br />

14.3 PLGA as a Polymer for Nanoparticles ............................................................................. 244<br />

14.4 Application of PLGA/PLA Nanoparticles as Drug Delivery Vehicles<br />

to Cancer Tissues .............................................................................................................. 245<br />

14.5 PLGA Nanoparticles for Gene Delivery to Cancer.......................................................... 247<br />

14.6 PLGA Nanoparticles for Photodynamic Therapy............................................................. 248<br />

14.7 Concluding Remarks ......................................................................................................... 248<br />

Acknowledgments ........................................................................................................................ 248<br />

References..................................................................................................................................... 249<br />

14.1 INTRODUCTION<br />

Despite significant efforts in the field of oncology, cancer remains one of the leading causes of death<br />

in industrialized countries. 1 Of the various options available, surgery and radiation therapy are<br />

commonly used to treat localized tumors whereas chemotherapy is primarily used in the management<br />

and elimination of hematological malignancies and metastasized tumors. 2 The major<br />

limitation of the current cancer chemotherapeutics is their high pharmacokinetic volume of distribution<br />

and rapid elimination rates, requiring more frequent and high dose administration of these<br />

agents. 3 Therefore, cancer chemotherapeutics cause unacceptable damage to normal tissue when<br />

used at doses required to eradicate cancer cells. Furthermore, multidrug resistance in tumor cells<br />

exacerbates the problem of cancer chemotherapy. 4–7 Selective delivery of anti-cancer drugs to the<br />

tumor tissue offers a formidable solution to the problem. This could significantly enhance the<br />

therapeutic efficacy of these drugs and diminish their toxicity. 8,9 The overall goal of drug delivery<br />

strategies is to selectively attack cancer cells while saving normal tissue from drug toxicity. 10,11<br />

The other issue with cancer chemotherapeutics is their systemic delivery, as most of these agents<br />

are poorly water soluble. Hence they require adjuvant or excipients to dissolve, which are often<br />

associated with serious side effects. 12 Therefore, drug carrier systems which would address the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

243


244<br />

above problem of delivery as well as target drugs selectively to the tumor would be a major<br />

advancement in cancer chemotherapy. Many macromolecular carriers, including soluble synthetic<br />

and natural polymers, implants, liposomes, microspheres, dendrimers, nanoparticles, etc., have<br />

been tested for selective delivery of drugs to the tumor tissue. 13<br />

14.2 GENERAL PRINCIPLES OF DRUG TARGETING TO CANCER<br />

14.2.1 PASSIVE TARGETING<br />

Passive targeting refers to the accumulation of drug or drug-carrier system at a particular site due to<br />

physicochemical or pharmacological factors. Permeability of the tumor vasculature increases to the<br />

point where particulate carriers such as nanoparticles can extravasate from blood circulation and<br />

localize in the tumor tissue. 14,15 This occurs because as tumors grow and begin to outstrip the<br />

available supply of oxygen and nutrients, they release cytokines and other signaling molecules that<br />

recruit new blood vessels to the tumor, a process known as angiogenesis. 16 Angiogenic blood<br />

vessels, unlike the tight blood vessels in most normal tissues, have gaps as large as 600–800 nm<br />

between adjacent endothelial cells. Drug carriers in the nanometer size range can extravasate<br />

through these gaps into the tumor interstitial space. 17,18 Because tumors have impaired lymphatic<br />

drainage, the carriers concentrate in the tumor, resulting in higher drug concentration in the tumor<br />

tissue (10-fold or higher) than that can be achieved with the same dose of free drug. This is<br />

commonly referred to as enhanced permeability and retention, or the EPR effect.<br />

14.2.2 ACTIVE TARGETING<br />

Active targeting to the tumor can be achieved <strong>by</strong> molecular recognition of cancer cells either via<br />

ligand–receptor or antibody–antigen interactions. Active targeting may also lead to receptormediated<br />

cell internalization of drug carrier system. Nanoparticles and other polymer drugconjugates<br />

offer numerous opportunities for targeting tumors through surface modifications<br />

which allow specific biochemical interactions with the proteins/receptorsexpressedontarget<br />

cells. 19,20 For active and passive targeting of drug carrier systems, it is essential to avoid their<br />

uptake <strong>by</strong> the reticuloendothelial system (RES) so that they remain in the blood circulation and<br />

extravasate in the tumor vasculature. Particles with more hydrophobic surfaces are preferentially<br />

taken up <strong>by</strong> the liver, followed <strong>by</strong> the spleen and lungs. 21,22 Size of nanoparticles as well as their<br />

surface characteristics are the key parameters that can alter the biodistribution of nanoparticles.<br />

Particles smaller than 100 nm and coated with hydrophilic polymers such as amphiphilic polymeric<br />

compounds which are made of polyethylene oxide such as poloxamers, poloxamines, or polyethylene<br />

glycol (PEG) are being investigated to avoid their uptake <strong>by</strong> the RES. To improve the efficacy<br />

of targeting cancer chemotherapeutics to the tumor, a combination of passive and active targeting<br />

strategy is being investigated where long-circulating drug carriers are conjugated to tumor cellspecific<br />

antibody or peptides. 23 In addition to the above approach, direct intratumoral injection of<br />

the carrier system is feasible if the tumor is localized and can be accessed for administration of a<br />

carrier system. 24<br />

14.3 PLGA AS A POLYMER FOR NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

A number of different polymers, both synthetic and natural, have been utilized in formulating<br />

biodegradable nanoparticles. Synthetic polymers have the advantage of sustaining the release of<br />

the encapsulated therapeutic agent over a period of days to several weeks as compared to natural<br />

polymers which, in general, have a relatively short duration of drug release. The polymers used for<br />

the formulation of nanoparticles include synthetic polymers such as polylactide–polyglycolide<br />

copolymers, polyacrylates, and polycaprolactones, or natural polymers such as albumin, gelatin,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Biodegradable PLGA/PLA Nanoparticles for Anti-Cancer Therapy 245<br />

alginate, collagen, and chitosan. Of these polymers, polylactides (PLA) and poly (D,L-lactide-coglycolide)<br />

(PLGA) have been most extensively investigated for drug delivery applications. 19<br />

PLGA/PLA-based polymers have a number of advantages over other polymers used in drug and<br />

gene delivery, such as their biodegradability, biocompatibility, and approval <strong>by</strong> the FDA for human<br />

use. 19,25 PLGA/PLA polymers degrade in the body through hydrolytic cleavage of the ester linkage<br />

to lactic and glycolic acid, although there are reports of involvement of enzymes in their biodegradation.<br />

19,25 These monomers are easily metabolized in the body via Krebs’ cycle and eliminated<br />

as carbon dioxide and water. Biodegradation products of PLGA and PLA polymers are formed at a<br />

very slow rate, and they therefore do not affect normal cell function. 26 Furthermore, these polymers<br />

have been tested for toxicity and safety in extensive animal studies and are currently used in<br />

humans for resorbable sutures, bone implants and screws, contraceptive implants, 27,28 and also<br />

as graft materials for artificial organs and supporting scaffolds in tissue engineering research. 28–30<br />

Long-term biocompatibility of these polymers was demonstrated <strong>by</strong> the absence of any untoward<br />

effects on intravascular administration of nanoparticles formulated using these polymers to the<br />

arterial tissue in pig and rat models of restenosis. 31<br />

14.4 APPLICATION OF PLGA/PLA NANOPARTICLES AS DRUG DELIVERY<br />

VEHICLES TO CANCER TISSUES<br />

There are several studies regarding PLGA/PLA nanoparticles or some modification of these polymers<br />

for delivery of anti-cancer agents and other therapeutic agents. 19 We have recently<br />

demonstrated increased efficacy of transferrin conjugated paclitaxel-loaded PLGA nanoparticles<br />

(Figure 14.1a and b) both in vitro and in an animal model of prostate carcinoma. 24 Transferrin<br />

receptors are over-expressed in most cancer cells <strong>by</strong> two to tenfold more than in normal cells. We<br />

have demonstrated that transferring-conjugated nanoparticles have enhanced cellular uptake<br />

(Figure 14.1c) and retention than unconjugated nanoparticles. 4 A single-dose intratumoral injection<br />

of transferrin conjugated nanoparticles in animal models of prostate carcinoma demonstrated<br />

complete tumor regression and higher survival rate than animals that received either drug in<br />

solution or unconjugated nanoparticles. 24 The IC50 for paclitaxel with transferrin conjugated nanoparticles<br />

was fivefold lower than that with unconjugated nanoparticles or with drug in solution in<br />

PC3 24 (Figure 14.1d) and in MCF-7 cells. 4 Kim et al. have demonstrated enhanced intracellular<br />

delivery of PLGA nanoparticles, which were surface-coated with cationic di-block copolymer,<br />

poly(L-lysine)–poly(ethylene glycol)–folate (PLL–PEG–FOL), in KB cells that overexpress<br />

folate receptors. 32 In another study, paclitaxel-loaded PLGA nanoparticles, which were conjugated<br />

to wheat germ agglutinin (WGA), demonstrated greater anti-proliferation activity in A549 and<br />

H1299 cells as compared to the conventional paclitaxel formulations. This enhanced activity of<br />

WGA-conjugated nanoparticles was attributed to greater intracellular accumulation of drug via<br />

WGA-receptor-mediated endocytosis of conjugated nanoparticles. 33 Cegnar et al. have developed<br />

cystatin-loaded PLGA nanoparticles with the strategy of inhibiting the tumor-associated activity of<br />

intracellular cysteine proteases cathepsins B and L. In an in vitro study, cystatin-loaded PLGA<br />

nanoparticles demonstrated 160-fold greater cytotoxic effect in MCF-10A neoT cells than free<br />

cystatin. 34 Similarly, interferon-alpha (IFN-alpha) loaded PLGA nanoparticles are being developed<br />

to improve the therapeutic efficacy of IFN-alpha while reducing its dose-related side effects. 35 In<br />

studies <strong>by</strong> Yoo et al., doxorubicin was chemically conjugated to a terminal end group of PLGA <strong>by</strong><br />

an ester linkage and the doxorubicin–PLGA conjugate was formulated into nanoparticles. Nanoparticles<br />

containing the conjugate exhibited sustained doxorubicin release profiles over a onemonth<br />

period, whereas those containing unconjugated doxorubicin showed a rapid drug release<br />

within five days. The conjugated doxorubicin nanoparticles demonstrated increased drug uptake in<br />

HepG2 cell line and exhibited slightly lower IC 50 value compared to that of free doxorubicin.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


246<br />

Uptake (μg/mg cell protein)<br />

(a)<br />

(c)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

NP NP-Tf NP-Tf+freeTf<br />

Treatments<br />

Cumulative % release<br />

% Growth<br />

0<br />

0 10 20 30 40 50 60 70<br />

The in vivo anti-tumor activity assay showed that a single injection of these nanoparticles had<br />

comparable activity to that of free doxorubicin when administered <strong>by</strong> daily injection. 36 Thus<br />

different strategies are being investigated using biodegradable PLGA-based nanoparticles to<br />

deliver anti-cancer drugs more effectively, both at the cellular level and also to the tumor tissue.<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

(b)<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

(d)<br />

Nanotechnology for Cancer Therapy<br />

Time (Day)<br />

1 10 100 1000<br />

Paclitaxel concentration (ng/ml)<br />

FIGURE 14.1 (a) Transmission electron micrographs of paclitaxel-loaded PLA nanoparticles (NPs) (barZ<br />

100 nm). (b) Cumulative release of paclitaxel (Tx) from NPs in vitro. Tx loading in NPs is 5.4% (w/w). Data as<br />

mean (Gs.e.m. (nZ3). (c) Uptake of unconjugated (NPs) and conjugated (NPs–Tf) in PC3 cells. A suspension<br />

of NPs (100 mg/mL) was incubated with PC3 cells (5!10 4 cells) for 1 h, cells were washed, and the NP levels<br />

in cells were determined <strong>by</strong> HPLC. To determine the competitive inhibition of uptake of Tf-conjugated NPs,<br />

excess dose of free Tf (50 mg) was added in the medium prior to incubation with NPs–Tf. Data as mean G<br />

s.e.m. (nZ6).* p!.05 NPs–TfC free Tf vs. NPs, ** p!.005 NPs–Tf vs. NPs. NPs contain a fluorescent<br />

marker, 6-coumarin, to quantify their uptake. (d) Dose-dependent cytotoxicity of paclitaxel (Tx) in PC3 cells.<br />

Different concentration of Tx either as solution (%) or encapsulated in NPs (Tx–NPs (&) or Tx–NPs–Tf, (:)<br />

was added to wells with NPs (without drug) or medium acting as respective controls. The medium was changed<br />

at two days and then on every alternate day with no further dose of the drug added. The extent of growth<br />

inhibition was measured at five days. The percentage of survival was determined <strong>by</strong> standardizing the absorbance<br />

of controls to 100% (nZ6). Data as mean Gs.e.m. * p!.005 Tx–NPs–Tf vs. Tx–sol and Tx–NPs. (Data<br />

reproduced from Sahoo, S. K. and Labhasetwar, V., Mol. Pharm., 2, 373, 2005; Sahoo, S. K., Ma, W., and<br />

Labhasetwar, V., Int. J. Cancer, 112, 335, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Biodegradable PLGA/PLA Nanoparticles for Anti-Cancer Therapy 247<br />

14.5 PLGA NANOPARTICLES FOR GENE DELIVERY TO CANCER<br />

Gene therapy to cancer represents one of the most rapidly developing areas in pre-clinical and<br />

clinical cancer research. Crucial to the success of any gene therapy strategy is the efficiency with<br />

which the gene is delivered. 37 This in turn is dependent upon the type of delivery vector used. Many<br />

vectors have been developed based on either recombinant viruses or non-viral vectors. Research<br />

utilizing viral vectors has progressed much more rapidly than that with non-viral vectors. This is<br />

reflected in the fact that approximately 85% of current clinical protocols involve viral vectors.<br />

However, these vectors are still limited in many ways, particularly in relation to issues of safety,<br />

immunogenicity, limitations on the size of the gene that can be delivered, specificity, production<br />

problems, toxicity, cost and others. 38,39<br />

Although various polymeric systems are under investigation, 40 our efforts are focused on<br />

investigating biodegradable nanoparticles as a gene delivery system in cancer therapy. 19,41–43<br />

These nanoparticles are formulated using PLGA and PLA polymers, with plasmid DNA (pDNA)<br />

entrapped into the nanoparticle matrix. The main advantage of PLGA or PLA-based nanoparticles<br />

for gene delivery is their non-toxic nature. Furthermore, the slow release of the encapsulated DNA<br />

from nanoparticles is expected to provide sustained gene delivery (Figure 14.2a). Blends of PLGA<br />

and polyoxyethylene derivatives such as poloxamer and poloxamine are also being used to<br />

modulate DNA release from nanoparticles. 44 In our study we demonstrated anti-proliferative<br />

activity of wild-type (wt) p53 gene-loaded nanoparticles in breast cancer cell line. 45 Cells transfected<br />

with wt-p53 DNA-loaded nanoparticles demonstrated sustained and significantly greater<br />

anti-proliferative effect than those treated with naked wt-p53 DNA (Figure 14.2b) or wt-p53<br />

DNA complexed with a commercially available transfecting agent (Lipofectamine). This effect<br />

was attributed to sustained nanoparticles-mediated gene expression. This was evident from<br />

sustained p53 mRNA levels observed in cells transfected with nanoparticles compared with<br />

levels in cells which were transfected with naked wt-p53 DNA.<br />

Fluorescence intensity<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Nanoparticles<br />

DNA<br />

% Proliferation<br />

0<br />

0.08 1 3 5 7<br />

(a) Time (day)<br />

(b)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

p53 DNA<br />

p53-Nanoparticles<br />

0 2 4 6 8<br />

Time (day)<br />

FIGURE 14.2 (a) Quantitative determination of intracellular DNA levels. Cells transfected with YOYOlabeled<br />

DNA-loaded nanoparticles demonstrated sustained and increase in intracellular DNA levels as<br />

opposed to transient DNA levels in the cells transfected with naked DNA. Data represented as mean G<br />

s.e.m., nZ6, * p!.001. (b) Anti-proliferative activity of wt-p53 DNA-loaded nanoparticles (NPs) and<br />

naked wt-p53 DNA (DNA) in MDAMB-435S cells. Cells (2500 cells/well) grown in a 96-well plate were<br />

incubated with DNA-loaded NPs or equivalent amount of naked DNA Cell growth was followed using a<br />

standard MTS assay. NPs demonstrated increase in anti-proliferative activity with incubation time. Data<br />

represented as meanGs.e.m., nZ6, * p!.01. (Data reproduced from Prabha, S. and Labhasetwar, V., Mol.<br />

Pharm., 1, 211, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


248<br />

Recently, He et al. have formulated thymidine kinase gene (TK gene)-loaded nanoparticles and<br />

investigated the expression of TK gene in hepatocarcinoma cells. 46 The expression of DNA<br />

encapsulated in nanoparticles was significantly greater than that with naked DNA. In another<br />

study, Cohen et al. have demonstrated significantly greater gene transfection (1–2 orders of magnitude)<br />

with PLGA nanoparticles than that with a liposomal formulation following their<br />

intramuscular injection in rats. Furthermore, the nanoparticle-mediated gene transfection was<br />

seen up to 28 days in the above study. Recently Kumar et al. have prepared cationic PLGA<br />

nanoparticles modified with cationic chitosan and demonstrated gene expression in A549 epithelial<br />

cells. 47<br />

14.6 PLGA NANOPARTICLES FOR PHOTODYNAMIC THERAPY<br />

Photodynamic therapy (PDT) is a promising new treatment modality 48 which involves administration<br />

of a photosensitizing drug. 49 Upon illumination at a suitable wavelength, it induces a<br />

photochemical reaction resulting in generation of reactive oxygen species (ROS) that can kill<br />

tumor cells directly as well as the tumor-associated vasculature, leading to tumor infarction.<br />

Targeting is essential in PDT as singlet oxygen is extremely reactive. Several strategies such as<br />

chemical conjugation with various water-soluble polymers and encapsulation into colloidal carriers<br />

have been considered for the parenteral administration of photosensitizers. These colloidal carriers<br />

include liposomes, emulsions, polymeric micelles and nanoparticles. PLGA nanoparticles were<br />

evaluated for PDT using meso-tetra(4-hydroxyphenyl) porphyrin as photosensitizer. 50 Vargas et al.<br />

have encapsulated porphyrin in PLGA nanoparticles and demonstrated enhanced photodynamic<br />

activity against mammary tumor cells than free drug. 51 Colloidal carrier associated with photosensitizer<br />

showed enhanced photodynamic efficiency and selectivity of tumor targeting as<br />

compared with dye administered in homogenous aqueous solution. 52,53 Recently, Saxena et al.<br />

have formulated PLGA nanoparticles loaded with indocyanine green (ICG) and determined their<br />

biodistribution in mice. The results demonstrated that the nanoparticle formulation significantly<br />

increased the ICG concentration and circulation time in plasma as well as the ICG uptake, accumulation<br />

and retention in various organs as compared to ICG solution. 54 Such a formulation can be<br />

explored in tumor-diagnosis as well as for PDT.<br />

14.7 CONCLUDING REMARKS<br />

It is essential to understand the molecular mechanisms involved in nanoparticle-mediated drug or<br />

gene delivery to explore their therapeutic potentials for cancer therapy. Also, the important questions<br />

we need to address are: What are the barriers in targeted cancer drug therapy? What is the<br />

efficiency of drug targeting with carrier systems? Can it provide a therapeutic dose of the drug in the<br />

tumor tissue? How long should the drug effect in the tumor tissue be sustained to regress it<br />

completely? Is a combination of drugs, especially those that work <strong>by</strong> different pathways, more<br />

effective than single-drug therapy? These and other questions are critical as we move from a<br />

conceptual stage to reality. With better understanding of molecular targets, discovery of more<br />

potent drugs and simultaneous developments in nanotechnology it seems that an effective cancer<br />

therapy is forthcoming.<br />

ACKNOWLEDGMENTS<br />

Nanotechnology for Cancer Therapy<br />

Grant support from the National Institutes of Health (R01 EB003975) is gratefully acknowledged.<br />

The authors also thank Ms. Elaine Payne and Ms. DeAnna Loibl for providing<br />

administrative support.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Biodegradable PLGA/PLA Nanoparticles for Anti-Cancer Therapy 249<br />

REFERENCES<br />

1. Greish, K. et al., Macromolecular therapeutics: Advantages and prospects with special emphasis on<br />

solid tumour targeting, Clin. Pharmacokinet., 42, 1089, 2003.<br />

2. Luo, Y. and Prestwich, G. D., Cancer-targeted polymeric drugs, Curr. Cancer Drug Targets, 2, 209,<br />

2002.<br />

3. Fang, J., Sawa, T., and Maeda, H., Factors and mechanism of “EPR” effect and the enhanced antitumor<br />

effects of macromolecular drugs including SMANCS, Adv. Exp. Med. Biol., 519, 29, 2003.<br />

4. Sahoo, S. K. and Labhasetwar, V., Enhanced antiproliferative activity of transferrin-conjugated<br />

paclitaxel-loaded nanoparticles is mediated via sustained intracellular drug retention, Mol. Pharm.,<br />

2, 373, 2005.<br />

5. Saeki, T. et al., Drug resistance in chemotherapy for breast cancer, Cancer Chemother. Pharmacol.,<br />

56(Suppl 7), 84, 2005.<br />

6. Kabanov, A. V., Batrakova, E. V., and Miller, D. W., Pluronic block copolymers as modulators of<br />

drug efflux transporter activity in the blood–brain barrier, Adv. Drug Deliv. Rev., 55, 151, 2003.<br />

7. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., Pluronic block copolymers for overcoming<br />

drug resistance in cancer, Adv. Drug Deliv. Rev., 54, 759, 2002.<br />

8. Jain, K. K., Nanotechnology-based drug delivery for cancer, Technol. Cancer Res. Treat., 4, 407,<br />

2005.<br />

9. Singh, K. K., Nanotechnology in cancer detection and treatment, Technol. Cancer Res. Treat., 4, 583,<br />

2005.<br />

10. Meisheid, A. M., Targeted therapies in the treatment of cancer, J. Contin. Educ. Nurs., 36, 193, 2005.<br />

11. Reddy, L. H., Drug delivery to tumours: Recent strategies, J. Pharm. Pharmacol., 57, 1231, 2005.<br />

12. Adams, J. D. et al., Taxol: a history of pharmaceutical development and current pharmaceutical<br />

concerns, J. Natl. Cancer Inst. Monogr., 15, 141, 1993.<br />

13. Sahoo, S. K. and Labhasetwar, V., Nanotech approaches to drug delivery and imaging, Drug Discov.<br />

Today, 8, 1112, 2003.<br />

14. Maeda, H. et al., Tumor vascular permeability and the EPR effect in macromolecular therapeutics:<br />

a review, J. Control. Release, 65, 271, 2000.<br />

15. Maeda, H., Sawa, T., and Konno, T., Mechanism of tumor-targeted delivery of macromolecular drugs,<br />

including the EPR effect in solid tumor and clinical overview of the prototype polymeric drug<br />

SMANCS, J. Control. Release, 74, 47, 2001.<br />

16. Folkman, J. and Shing, Y., Angiogenesis, J. Biol. Chem., 267, 10931, 1992.<br />

17. Jain, R. K., Integrative pathophysiology of solid tumors: Role in detection and treatment, Cancer<br />

J. Sci. Am., 4(Suppl 1), S48, 1998.<br />

18. Jain, R. K., Delivery of molecular and cellular medicine to solid tumors, J. Control. Release, 53, 49,<br />

1998.<br />

19. Panyam, J. and Labhasetwar, V., Biodegradable nanoparticles for drug and gene delivery to cells and<br />

tissue, Adv. Drug Deliv. Rev., 55, 329, 2003.<br />

20. Minko, T. et al., Molecular targeting of drug delivery systems to cancer, Curr. Drug Targets, 5, 389,<br />

2004.<br />

21. Gref, R. et al., Biodegradable long-circulating polymeric nanospheres, Science., 263, 1600, 1994.<br />

22. Gref, R. et al., Poly(ethylene glycol)-coated nanospheres: Potential carriers for intravenous drug<br />

administration, Pharm. Biotechnol., 10, 167, 1997.<br />

23. Vasir, J. K. and Labhasetwar, V., Targeted drug delivery in cancer therapy, Technol. Cancer Res.<br />

Treat., 4, 363, 2005.<br />

24. Sahoo, S. K., Ma, W., and Labhasetwar, V., Efficacy of transferrin-conjugated paclitaxel-loaded<br />

nanoparticles in a murine model of prostate cancer, Int. J. Cancer, 112, 335, 2004.<br />

25. Jain, R. A., The manufacturing techniques of various drug loaded biodegradable poly(lactide-coglycolide)<br />

(PLGA) devices, Biomaterials, 21, 2475, 2000.<br />

26. Shive, M. S. and Anderson, J. M., Biodegradation and biocompatibility of PLA and PLGA microspheres,<br />

Adv. Drug Deliv. Rev., 28, 5, 1997.<br />

27. Hanafusa, S. et al., Biodegradable plate fixation of rabbit femoral shaft osteotomies. A comparative<br />

study, Clin. Orthop. Relat. Res., 262, 1995.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


250<br />

Nanotechnology for Cancer Therapy<br />

28. Matsusue, Y. et al., Tissue reaction of bioabsorbable ultra high strength poly (L-lactide) rod. A longterm<br />

study in rabbits, Clin. Orthop. Relat. Res., 246, 1995.<br />

29. Langer, R., Tissue engineering: a new field and its challenges, Pharm. Res., 14, 840, 1997.<br />

30. Mooney, D. J. et al., Long-term engraftment of hepatocytes transplanted on biodegradable polymer<br />

sponges, J. Biomed. Mater. Res., 37, 413, 1997.<br />

31. Guzman, L. A. et al., Local intraluminal infusion of biodegradable polymeric nanoparticles. A novel<br />

approach for prolonged drug delivery after balloon angioplasty, Circulation, 94, 1441, 1996.<br />

32. Kim, S. H. et al., Target-specific cellular uptake of PLGA nanoparticles coated with poly(L-lysine)–<br />

poly(ethylene glycol)–folate conjugate, Langmuir, 21, 8852, 2005.<br />

33. Mo, Y. and Lim, L. Y., Paclitaxel-loaded PLGA nanoparticles: Potentiation of anticancer activity <strong>by</strong><br />

surface conjugation with wheat germ agglutinin, J. Control. Release, 108, 244, 2005.<br />

34. Cegnar, M. et al., Poly(lactide-co-glycolide) nanoparticles as a carrier system for delivering cysteine<br />

protease inhibitor cystatin into tumor cells, Exp. Cell Res., 301, 223, 2004.<br />

35. Sanchez, A. et al., Biodegradable micro- and nanoparticles as long-term delivery vehicles for<br />

interferon-alpha, Eur. J. Pharm. Sci., 18, 221, 2003.<br />

36. Yoo, H. S. et al., In vitro and in vivo anti-tumor activities of nanoparticles based on doxorubicin–<br />

PLGA conjugates, J. Control. Release, 68, 419, 2000.<br />

37. Yamamoto, M. and Curiel, D. T., Cancer gene therapy, Technol. Cancer Res. Treat., 4, 315, 2005.<br />

38. Seth, P., Vector-mediated cancer gene therapy: an overview, Cancer Biol. Ther., 4, 512, 2005.<br />

39. Maitland, N. J., Stanbridge, L. J., and Dussupt, V., Targeting gene therapy for prostate cancer, Curr.<br />

Pharm. Des., 10, 531, 2004.<br />

40. Vasir, J. K. and Labhasetwar, V., Polymeric nanoparticles for gene delivery, Expt. Opin. Drug Del.,<br />

3, 325, <strong>2006</strong>.<br />

41. Panyam, J. et al., Rapid endo-lysosomal escape of poly(DL-lactide-co-glycolide) nanoparticles:<br />

Implications for drug and gene delivery, FASEB J., 16, 1217, 2002.<br />

42. Prabha, S. et al., Size-dependency of nanoparticle-mediated gene transfection: Studies with fractionated<br />

nanoparticles, Int. J. Pharm., 244, 105, 2002.<br />

43. Prabha, S. and Labhasetwar, V., Critical determinants in PLGA/PLA nanoparticle-mediated gene<br />

expression, Pharm. Res., 21, 354, 2004.<br />

44. Csaba, N. et al., PLGA: Poloxamer and PLGA:Poloxamine blend nanoparticles: New carriers for gene<br />

delivery, Biomacromolecules, 6, 271, 2005.<br />

45. Prabha, S. and Labhasetwar, V., Nanoparticle-mediated wild-type p53 gene delivery results in<br />

sustained antiproliferative activity in breast cancer cells, Mol. Pharm., 1, 211, 2004.<br />

46. He, Q. et al., Preparation and characteristics of DNA-nanoparticles targeting to hepatocarcinoma<br />

cells, World J. Gastroenterol., 10, 660, 2004.<br />

47. Kumar, M. N. et al., Cationic poly(lactide-co-glycolide) nanoparticles as efficient in vivo gene transfection<br />

agents, J Nanosci. Nanotechnol., 4, 990, 2004.<br />

48. Dolmans, D. E., Fukumura, D., and Jain, R. K., Photodynamic therapy for cancer, Nat. Rev. Cancer,3,<br />

380, 2003.<br />

49. Huang, Z., A review of progress in clinical photodynamic therapy, Technol. Cancer Res. Treat., 4,<br />

283, 2005.<br />

50. Konan, Y. N. et al., Encapsulation of p-THPP into nanoparticles: Cellular uptake, subcellular<br />

localization and effect of serum on photodynamic activity, Photochem. Photobiol., 77, 638, 2003.<br />

51. Vargas, A. et al., Improved photodynamic activity of porphyrin loaded into nanoparticles: an in vivo<br />

evaluation using chick embryos, Int. J. Pharm., 286, 131, 2004.<br />

52. Konan, Y. N. et al., Enhanced photodynamic activity of meso-tetra(4-hydroxyphenyl)porphyrin <strong>by</strong><br />

incorporation into sub-200 nm nanoparticles, Eur. J. Pharm. Sci., 18, 241, 2003.<br />

53. Konan, Y. N. et al., Preparation and characterization of sterile sub-200 nm meso-tetra (4-hydroxylphenyl)porphyrin-loaded<br />

nanoparticles for photodynamic therapy, Eur. J. Pharm. Biopharm., 55, 115,<br />

2003.<br />

54. Saxena, V., Sadoqi, M., and Shao, J., Polymeric nanoparticulate delivery system for Indocyanine<br />

green: Biodistribution in healthy mice, Int. J. Pharm., 308, 200, <strong>2006</strong>.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


15<br />

CONTENTS<br />

Poly(Alkyl Cyanoacrylate)<br />

Nanoparticles for Delivery<br />

of Anti-Cancer Drugs<br />

R. S. R. Murthy and L. Harivardhan Reddy<br />

15.1 Cancer Therapy ................................................................................................................. 252<br />

15.2 Nanoparticles as Drug Delivery Systems ......................................................................... 252<br />

15.2.1 Polymeric Nanoparticles ..................................................................................... 253<br />

15.2.1.1 Nanoparticles Prepared <strong>by</strong> the Polymerization Process .................... 253<br />

15.2.1.1.1 Dispersion Polymerization .............................................. 253<br />

15.2.1.1.1 Emulsion Polymerization ................................................ 254<br />

15.2.1.2 Poly(Alkyl Cyanoacrylate) Nanoparticles ......................................... 254<br />

15.2.1.3 Poly(Alkyl Cyanoacrylate) Nanocapsules ......................................... 256<br />

15.2.2 Loading of Drugs to Poly(Alkyl Cyanoacrylate) Nanoparticles........................ 256<br />

15.2.3 Characterization of Nanoparticles....................................................................... 258<br />

15.2.3.1 Physicochemical Characterization ..................................................... 258<br />

15.2.3.2 Degradation of Poly(Alkyl Cyanoacrylate) Nanoparticles................ 260<br />

15.2.3.3 Influence of Enzymes on the Stability<br />

of Poly(Alkyl Cyanoacrylate) Nanoparticles..................................... 260<br />

15.2.3.4 Drug Release from Poly(Alkyl Cyanoacrylate) Nanoparticles ......... 261<br />

15.2.4 In Vivo Distribution of Poly(Alkyl Cyanoacrylate) Nanoparticles ................... 261<br />

15.2.4.1 Surface Modification to Alter Biodistribution................................... 264<br />

15.2.5 Toxicity of Polyalkyl Cyanoacrylate Nanoparticles........................................... 264<br />

15.2.5.1 Toxicity of Drugs Associated with Poly(Alkyl Cyanoacrylate)<br />

Nanoparticles ...................................................................................... 266<br />

15.2.6 Drug Delivery in Cancer with Poly(Alkyl Cyanoacrylate) Nanoparticles ........ 266<br />

15.2.6.1 Targeting to Cancer Cells and Tissues .............................................. 266<br />

15.2.6.2 Drug Delivery to Resistant Tumor Cells ........................................... 268<br />

15.2.6.3 Drug Delivery to Hepatocellular Carcinomas ................................... 272<br />

15.2.6.4 Drug Delivery to Brain Tumors......................................................... 274<br />

15.2.6.5 Drug Delivery to Breast Cancer ........................................................ 275<br />

15.2.6.6 Drug Delivery to Lymphatic Carcinomas ......................................... 276<br />

15.2.6.7 Drug Delivery to Gastric Carcinomas ............................................... 276<br />

15.2.6.8 Chemoembolization Using Alkylcyanoacrylates............................... 277<br />

15.2.6.9 Delivery of Peptide Anti-Cancer Drugs ............................................ 278<br />

15.3 Conclusions ....................................................................................................................... 280<br />

References .................................................................................................................................... 280<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

251


252<br />

15.1 CANCER THERAPY<br />

Cancer is defined <strong>by</strong> two characteristics: uncontrolled cell growth and the ability to invade, metastasize,<br />

and spread to distant sites. Despite several modes of therapy—such as chemotherapy,<br />

immunotherapy, and radiotherapy—the therapy of cancer remains a challenge. Systemic<br />

chemotherapy is the widely adopted treatment for the disseminated malignant disease, while<br />

recent progress in drug therapy has resulted in curative chemotherapeutic regimens for several<br />

tumors. However, chronic administration of anti-cancer drugs leads to severe systemic toxicity. To<br />

improve the therapy with these anti-cancer agents and reduce this associated toxicity, drug delivery<br />

systems have been introduced to deliver the drugs directly to the site of interest. 1 Normal cells are<br />

also susceptible to the cytotoxic effects of chemotherapeutic agents and exhibit a dose response<br />

effect, but the response curve is shifted relative to that of malignant cells. Proliferative normal<br />

tissues such as bone marrow and gastrointestinal mucosa are generally the most susceptible<br />

to chemotherapy-induced toxicity. Because chemotherapy has been widely adopted for the treatment<br />

of cancer, several chemotherapeutic agents with potential anti-cancer activity have<br />

been introduced.<br />

Several drug delivery systems facilitate effective chemotherapy with the anti-cancer agents.<br />

These systems include liposomes, microparticles, supramolecular biovectors, polymeric micelles,<br />

and nanoparticles. Although research on targeted therapeutic systems for cancer therapy has not<br />

significantly contributed to their commercialization for human application, the introduction into the<br />

market of products like doxorubicin long-circulating liposomes (Doxil) and Polifeprosan-20 with<br />

carmustine (BCNU, Gliadel Wafer) for cancer therapy has renewed interest in the field of targeted<br />

drug delivery to cancers. 2<br />

15.2 NANOPARTICLES AS DRUG DELIVERY SYSTEMS<br />

Nanotechnology for Cancer Therapy<br />

The challenge of modern drug therapy is the optimization of the pharmacological action of the<br />

drugs coupled with the reduction of their toxic effects in vivo. The prime objectives in the design of<br />

drug delivery systems are the controlled delivery of the drug to its site of action at a therapeutically<br />

optimal rate and dosage to avoid toxicity and improve the drug effectiveness and therapeutic index.<br />

Among the most promising systems to achieve this goal are the colloidal drug delivery systems,<br />

which include liposomes, niosomes, microemulsions, and nanoparticles. Nanoparticles are<br />

considered promising colloidal drug carriers, as they overcome the technological limitations and<br />

stability problems associated with liposomes, niosomes, and microemulsions. For instance,<br />

camptothecin-based drugs, because of their poor solubility and labile lactone ring, pose challenges<br />

for drug delivery. Williams et al. 3 developed a nanoparticle delivery system for a camptotheca<br />

alkaloid (SN-38) that is stable in human serum albumin; high lactone concentrations were observed<br />

even after 3 h and showed prolonged in vivo half-life of the active (lactone) form.<br />

Nanoparticles are solid colloidal particles ranging in size from 10 to 1000 nm. Depending on<br />

the process used for their preparation, two types of nanoparticles are obtained: nanospheres and<br />

nanocapsules (Figure 15.1). Nanospheres possess a rigid matrix structure that incorporates the drug,<br />

whereas the nanocapsule contains an oily core incorporating the drug surrounded <strong>by</strong> a membrane<br />

structure. 4 These are prepared from either synthetic or natural macromolecules in which the drug is<br />

dissolved, entrapped, encapsulated, or adsorbed on the surface. Generally, the definition of nanoparticles<br />

includes not only the particles described <strong>by</strong> Birrenbach and Speiser 5 <strong>by</strong> the term<br />

“Nanopellets,” but also nanocapsules and polymer lattices such as the “molecular scale drug<br />

entrapment” products described <strong>by</strong> Rhodes et al. 6 and Boylan and Banker. 7 Nanoparticles<br />

possess a matrix structure that incorporates the pharmacologically active substance and facilitates<br />

the controlled release of the active agent.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 253<br />

Polymeric matrix<br />

containing drug<br />

Nanosphere<br />

15.2.1 POLYMERIC NANOPARTICLES<br />

In recent years, biodegradable polymeric nanoparticles have attracted attention as potential drug<br />

carriers because of their applications in the controlled delivery of drugs, their ability to target<br />

organs and tissues, their function as carriers for DNA in gene therapy, and their ability to perorally<br />

deliver proteins, peptides, and genes. 8,9 Langer 8 stated that the construction of fully biomimetic<br />

matrices containing all of the required chemical and topographical factors has not yet been<br />

achieved. Scientists are focusing on the design of new polymeric materials in which both topographical<br />

features and chemical cues act synergistically to qualify as drug-delivery carriers.<br />

Polymeric nanoparticles can be prepared either <strong>by</strong> the polymerization of monomers or from the<br />

preformed polymers. This chapter will focus on the former process.<br />

15.2.1.1 Nanoparticles Prepared <strong>by</strong> the Polymerization Process<br />

Historically, the first methods used to produce nanoparticles were derived from the field of latex<br />

engineering developed <strong>by</strong> polymer chemists. These methods were based on the in situ polymerization<br />

of monomers in various media. In the early 1970s, Birrenbach and Speiser, the<br />

pioneers in the field, produced the first polymerized nanoparticles for pharmaceutical use. Two<br />

types of polymerization processes were adopted for the preparation of polymeric nanoparticles:<br />

dispersion polymerization (DP) and emulsion polymerization (EP).<br />

15.2.1.1.1 Dispersion Polymerization<br />

Oily core<br />

+ Drug<br />

Nanocapsule<br />

FIGURE 15.1 Representation of nanosphere and nanocapsule.<br />

Polymeric membrane<br />

A DP involves a monomer, an initiator, a solvent in which the newly formed polymer is insoluble,<br />

and a polymeric stabilizer. The polymer is formed in the continuous phase and precipitates into a<br />

new particle phase stabilized <strong>by</strong> the polymeric stabilizer. The nucleation is directly induced in the<br />

aqueous monomer solution. Small particles are formed <strong>by</strong> the aggregation of growing polymer<br />

chains precipitating from the continuous phase when these chains exceed a critical chain length.<br />

When enough stabilizer covers the particles, coalescence of these precursor particles with themselves<br />

and with their aggregates results in the formation of stable colloidal particles. In this<br />

technique, the monomer is dispersed or dissolved in an aqueous medium from which the formed<br />

polymer precipitates.<br />

Essentially, the initiation of a monomer for polymerization requires an initiator that can<br />

generate ions or radicals to start the polymerization process. If the nucleation of the monomer is<br />

due to ions, then the mechanism is called “ionic polymerization.” Depending on the type of ions<br />

produced, the ionic polymerization may be anionic or cationic. If the radical nucleates the<br />

monomer, then the mechanism is known as “radical polymerization.” 10<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


254<br />

15.2.1.1.2 Emulsion Polymerization<br />

Emulsion polymerization is a method for producing a latex and a polymer that exhibit any desired<br />

morphology, composition, sequence distribution, surface groups, molecular weight distribution,<br />

particle concentration, or other characteristics. In this technique, the monomer is emulsified in a<br />

nonsolvent-containing surfactant, which leads to the formation of monomer-swollen micelles and<br />

stabilized monomer droplets. The polymerization is performed in the presence of an initiator. The<br />

initiator generates either radicals or ions depending on the type of initiator, and these radicals or<br />

ions nucleate the monomeric units and start the polymerization process. The monomer-swollen<br />

micelles exhibit sizes in the nanometer range and thus have a much larger surface area than the<br />

monomer droplets. It has been assumed that, once formed, the free reactive monomers would more<br />

probably initiate the reaction within the micelles itself. In this case, the monomer droplets would act<br />

as monomer reservoirs. Being slightly soluble in the surrounding phase, the monomer molecules<br />

reach the micelles <strong>by</strong> diffusion from the monomer droplets, thus allowing the polymerization to be<br />

continued in the micelles. 11,12 Generally, the reaction continues until the monomer completely<br />

disappears. The particles obtained <strong>by</strong> EP are usually smaller (100–300 nm) than the original<br />

stabilized monomer droplets in the continuous phase. EP may be performed using either organic<br />

or aqueous media as the continuous phase.<br />

Emulsion Polymerization in an Organic Continuous Phase. EP in the organic continuous phase<br />

was the first process reported for the preparation of nanoparticles. 5,13 In this process, the watersoluble<br />

monomers are polymerized. Acrylamide and the crosslinker N,N 0 -bisacrylamide were the<br />

prototype monomers to be polymerized using chemical initiators such as N,N,N 0 ,N 0 -tetramethylethylenediamine<br />

and potassium peroxdisulphate or <strong>by</strong> light irradiation <strong>by</strong> g- or UV-radiations.<br />

However, the high toxicity of monomers and the need for high amounts of organic solvents and<br />

surfactants limits the importance of this technique.<br />

The cyanoacrylate monomers are relatively less toxic than acrylamide, and their polymers are<br />

biodegradable. Poly(alkyl cyanoacrylate) (PACA) nanoparticles were prepared <strong>by</strong> EP in the<br />

continuous organic phase. In this case, the monomer was added to the organic phase, creating<br />

nanoparticles with a shell-like structure (nanocapsules), as well as solid particles. 14 The mechanism<br />

of this nanocapsule formation was explained as following. The drug dissolved in aqueous phase was<br />

solubilized in the organic phase (containing surfactants) such as iso-octane, cyclohexane–chloroform,<br />

isopropyl myristate–butanol, and hexane. This result in a microemulsion with water-swollen<br />

micelles containing the drug. The monomer diffuse into the swollen micelles, and the OH K ions<br />

initiated the polymerization. The polymerization is so rapid that only an impermeable wall could be<br />

formed at the organic/water interface, preventing the diffusion of further monomers into the<br />

particles. The high amounts of organic solvents and surfactants required for this process,<br />

however, has greatly limited its application.<br />

Emulsion Polymerization in an Aqueous Continuous Phase. This technique is widely used for<br />

the preparation of nanoparticles <strong>by</strong> the polymerization of a wide variety of monomers, including<br />

alkylcyanoacrylates. Employing very low quantities of surfactants, it is only used to stabilize the<br />

newly formed polymer particles. Apart from this, the polymerization of alkylcyanoacrylates has<br />

also been carried out in the absence of surfactants, 15 in aqueous media containing dextran or<br />

hydroxylpropyl-b-cyclodextrin (HPCD). Poly(ethyl cyanoacrylate) and poly(isobutyl cyanoacrylate)<br />

(PIBCA) nanospheres containing metaclopramide have been prepared using this technique.<br />

However, the resulting drug loading was only 9.2% and 14.8%, respectively.<br />

15.2.1.2 Poly(Alkyl Cyanoacrylate) Nanoparticles<br />

Nanotechnology for Cancer Therapy<br />

The first report on the synthesis of PACA nanoparticles appeared in 1979. 16 PACA nanoparticles<br />

are biodegradable and hence are eliminated rapidly from the body. 17 The cyanoacrylate monomers<br />

are polymerized in the aqueous medium <strong>by</strong> the anionic polymerization; 16,18 the mechanism is<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 255<br />

OH−<br />

+<br />

H2C<br />

N<br />

C<br />

C<br />

C<br />

O<br />

O<br />

R1<br />

HO<br />

H2C N<br />

C<br />

C<br />

C<br />

O<br />

O<br />

R1<br />

HO<br />

H2C<br />

depicted in Figure 15.2. The cyanoacrylate monomer is initiated <strong>by</strong> OH K ions, and the polymerization<br />

was performed at a low pH to control the reaction, which otherwise would have led<br />

to rapid polymerization resulting in the precipitation of large polymer aggregates. The stabilizers<br />

used were found to have a significant influence on the size of the formed polymer particles. High<br />

concentrations of poloxamer 188 (above 2%) reduced the particle size of polyisobutyl cyanoacrylate<br />

nanoparticles from around 200 nm without an emulsifier, to as low as 31–56 nm. 19<br />

Nanoparticles prepared without stabilizers or prepared using polysorbates were found to have<br />

monomodal molecular weight distribution, with mean molecular weights of 1000–4000 Da,<br />

whereas the stabilizers such as dextrans or poloxamers led to a distinctive bimodal distribution<br />

of the molecular weights, with peaks at 1000–4000 and 20,000–40,000 Da. This bimodal distribution<br />

is indicative of two separate polymerization reactions. The first reaction probably occurs in<br />

the aqueous phase where the termination of polymerization <strong>by</strong> H C ions is rapid, leading to small<br />

molecular weights. The second reaction is possibly the polymerization of captured growing unterminated<br />

polymer molecules within the primary particles. Due to the low concentration of H C ions<br />

in this environment, the termination frequency is reduced and hence the molecular weights would<br />

be much higher. Due to the initiation of cyanoacrylate polymerization <strong>by</strong> bases, the basic drugs also<br />

can act as initiators. However, because of the complexity of cyanoacrylate polymerization, the<br />

resulting particle size in such multi-component systems is difficult to predict.<br />

Poly(butyl cyanoacrylate) (PBCA) nanoparticles of n-butyl cyanoacrylate containing methotrexate<br />

were prepared <strong>by</strong> DP and EP. 20 DP nanoparticles were prepared using dextran as a<br />

stabilizer; the EP nanoparticles were stabilized <strong>by</strong> poloxamer 188. A high zeta potential was<br />

observed for nanoparticles prepared <strong>by</strong> the DP method, whereas the incorporation of methotrexate<br />

resulted in a decrease in zeta potential. The DP nanoparticles exhibited a high release of methotrexate,<br />

suggesting the channelizing effect of dextran chains incorporated into nanoparticles during<br />

polymerization. When tested in two different release media such as 0.1 mol L K1 HCl and pH 7.4<br />

phosphate buffer, a significant difference (p!0.01) in release rates was found for DP and EP<br />

nanoparticles. Drug release from both the nanoparticles followed Fickian diffusion in 0.1mol<br />

L K1 HCl, whereas the mechanism was found anomalous in pH 7.4 phosphate buffer.<br />

A similar study conducted on PBCA nanoparticles loaded with doxorubicin hydrochloride <strong>by</strong><br />

incorporation and adsorption techniques 21 showed rapid drug release in 0.001-N HCl from both<br />

DP and EP nanoparticles; the release kinetics followed the Higuchi equation.<br />

N<br />

C<br />

C−<br />

C<br />

O<br />

O<br />

R1<br />

H+ Monomer<br />

n H<br />

FIGURE 15.2 Mechanism of alkyl cyanoacrylate polymerization. (From Behan, N., Birkinshaw, C., and<br />

Clarke, N., Biomaterials, 22, 1335, 2001. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


256<br />

15.2.1.3 Poly(Alkyl Cyanoacrylate) Nanocapsules<br />

Nanocapsules are spherical structures formed <strong>by</strong> an envelope surrounding a liquid central cavity.<br />

The technique for the preparation of PIBCA nanocapsules was first reported <strong>by</strong> Fallaouh et al. 22<br />

The nanocapsules were obtained <strong>by</strong> injecting the organic phase, consisting of oil, with the dissolved<br />

drug, isobutylcyanoacrylate, and ethanol into the aqueous phase containing Pluronic F68 (a<br />

nonionic surfactant) under magnetic stirring. The colloidal suspension was concentrated <strong>by</strong> evaporation<br />

under a vacuum and filtered to obtain nanocapsules that possessed a shell-like structure<br />

with a wall 3-nm thick. This thickness was confirmed <strong>by</strong> transmission electron microscopy (TEM)<br />

after negative staining with phosphotungstic acid. The pH of the polymerization medium played an<br />

important role in obtaining nanocapsules. Nanocapsules with uniform size and low polydispersity<br />

were obtained only between pH 4 and 10. The concentration of alcohol above 15% only led to the<br />

formation of isolated nanocapsules. The mechanism of nanocapsule formation is probably the<br />

interfacial polymerization described <strong>by</strong> Florence et al. 23<br />

for the manufacture of<br />

PACA microcapsules.<br />

A different method was adopted <strong>by</strong> Chouinard et al. 24 for the preparation of poly(isohexyl<br />

cyanoacrylate) (PIHCA) nanocapsules. The monomer was treated with sulfur dioxide and dissolved<br />

in a solution containing Miglyol 810 (caprylic/capric triglyceride) in ethanol. This solution was<br />

added slowly, in a 1:2 ratio, into an aqueous phase containing a solution of 0.5% poloxamer 407 and<br />

10 mM phosphate buffer under stirring. The nanocapsules were then purified <strong>by</strong> centrifugation and<br />

washing with distilled water. In this method, the sulfur dioxide acts as an inhibitor, preventing the<br />

polymerization of the cyanoacrylates <strong>by</strong> the ethanol. With the addition of the oily phase to<br />

the aqueous phase, the sulfur dioxide and ethanol migrate to the external phase, resulting in the<br />

formation of a very fine emulsion. In this case, the nanocapsule size was mainly influenced <strong>by</strong><br />

Miglyol concentration. Freeze-fracture electron microscopy studies showed that polymeric walls of<br />

different thicknesses and densities can be prepared, which could be significant for the design of<br />

nanocapsules with tailored drug-release kinetics.<br />

Palumbo et al. 25 adopted a similar technique for the preparation of polyethyl-2-cyanoacrylate<br />

nanocapsules containing idebenone (an antioxidant). In this method, the acetonic solution of<br />

Miglyol 812, idebenone, and monomer was added to 100 mL of aqueous phase (pH 7) containing<br />

Tween 80. The presence of nonionic surfactant allowed the polymerization of the ethylcyanoacrylate<br />

at the oil–water interface, thus encapsulating Miglyol 812 droplets. The immediate<br />

polymerization triggered the formation of drug-loaded nanocapsules, with the suspensions concentrated<br />

under a vacuum to remove any acetone. Tween 80 reduced the hydrodynamic size of the<br />

emulsion droplets as a function of its concentration, resulting in smaller-sized nanocapsules. These<br />

nanocapsules exhibited a negative charge that was determined <strong>by</strong> zeta-potential measurements. The<br />

freeze-fracture electron microscopy revealed the presence of internal oil droplets surrounded <strong>by</strong> the<br />

polymeric shell of the nanocapsule.<br />

15.2.2 LOADING OF DRUGS TO POLY(ALKYL CYANOACRYLATE) NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

The drugs were loaded to PACA nanoparticles, either <strong>by</strong> incorporation during the polymerization<br />

process or <strong>by</strong> absorption onto the surface of the preformed particles. The former case can lead to the<br />

covalent coupling of the drug to the polymer, 26 or to the formation of a solid solution or solid<br />

dispersion of the drug in the polymer network. The addition of the drug to empty nanoparticles may<br />

lead to covalent coupling 27 of the drug; alternatively, the drug may be bound <strong>by</strong> sorption. The<br />

sorption can also lead either to the diffusion of the drug into the polymer network and the formation<br />

of a solid solution, 28 or to surface adsorption of the drug. 29 In the majority of cases, the type of drug<br />

loading will determine the carrier capacity. The carrier capacity is defined as the percentage of drug<br />

associated with a given amount of nanoparticles for a given initial concentration of drug without<br />

considering the characteristics of the adsorption isotherm. 30,31 Several drugs such as betaxolol<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 257<br />

chlorhydrate, 32 hematoporphyrin, 33 and primaquine 34 were adsorbed onto the preformed PACA<br />

nanoparticles, whereas drugs such as doxorubicin 35 and actinomycin D were incorporated into the<br />

nanoparticles during the polymerization process. The incorporation of rose bengal into the poly<br />

(butyl-2-cyanoacrylate) nanoparticles during the polymerization led to much greater carrier<br />

capacity than did the simple adsorption method. 29 The rose bengal incorporated into nanoparticles<br />

during polymerization is expected to be either dissolved, dispersed in, or adsorbed on the polymer<br />

matrix and remaining to be adsorbed on the particle surface. In contrast, after the adsorption<br />

process, the drug solely occupies the external surface of the nanoparticles, including minor pores<br />

and pits. Adsorption of hematoporphyrin onto PIBCA nanoparticles resulted in poor carrier<br />

capacity of the nanoparticles, because the drug was adsorbed mainly at the surface of the nanoparticles.<br />

33 A similar report <strong>by</strong> Beck et al. 28 describes the preparation of PBCA nanoparticles<br />

loaded with mitoxantrone using both the incorporation and adsorption methods. The proportion<br />

of mitoxantrone bound to the particles was analyzed to be about 15% of the initial drug concentration<br />

with the incorporation method and about 8% with the adsorption method.<br />

The type of drug binding onto the nanoparticles may result in different release mechanisms and<br />

rates. 36 Whether the drug is present in the form of a solid solution or a solid dispersion, its release<br />

characteristics mainly depend on the degradation rate of the polymer. 30 Drug desorption (tested in<br />

phosphate buffered medium pH 5.9) resulted essentially from a shift in the equilibrium conditions<br />

between free and adsorbed drug following the dilution of the nanoparticle suspension, modification<br />

of the pH conditions, or enzymatic hydrolysis <strong>by</strong> the added esterases. The release of hematoporphyrin<br />

was very rapid at physiological temperature and pH, and hence would not be useful for<br />

in vivo applications. 33 Finally, the surface charge of the particles and binding type between the drug<br />

and nanoparticles are the important parameters that determine the rate of desorption of the drug in<br />

the nanoparticles. 32 Recently, Poupaert and Couvreur 37 developed a computational derived structural<br />

model of doxorubicin interacting with oligomeric PACA in nanoparticles. The oligomeric<br />

PACAs are highly lipophilic entities and scavenge the amphiphilic doxorubicin during the polymerization<br />

process <strong>by</strong> extraction of protonated species from the aqueous environment to an<br />

increasingly lipophilic phase embodied <strong>by</strong> the growing PACAs. Interesting, hydrogen bonds<br />

were established between the N and H function of doxorubicin and the cyano groups of alkylcyanoacrylate.<br />

Therefore, the cohesion in this assembly comes from a blend of dipole-charge<br />

interaction, H-bonds, and hydrophobic forces. Upon hydrolytic erosion of the nanoparticles involving<br />

mainly the hydrolysis of ester groups to carboxylate, the assembly of doxorubicin–PACA<br />

tends to loosen the cohesion as the hydrophobic forces decline due to water intrusion and the<br />

increasing contribution of repulsive forces between anionic carboxylates. This creates, for<br />

example, a high local gradient of doxorubicin at the cell membrane surface, permitting MDR<br />

efflux to be overwhelmed. Very recently, heparin–PIBCA copolymer nanoparticles were developed<br />

as a carrier for hemoglobin, 38 which was indicated for treatment in thrombosis oxygen-deprived<br />

pathologies. In water, these copolymers spontaneously form nanoparticles with a ciliated surface of<br />

heparin. The heparin bound to the nanoparticle preserved its antithrombotic activity. The bound<br />

hemoglobin also maintained its capacity to bind ligands.<br />

The integration of drugs with polymerization additives was observed in the case of 5-fluorouracil<br />

(5-FU)-loaded PBCA nanoparticles. 39 5-FU in acidic solution (pH 2–3) may interfere in the<br />

initiation process through its amino groups via the formation of zwitterions. The proposed zwitterionic<br />

mechanism of initiation was supported <strong>by</strong> the molecular weight profiles of the polymer, as<br />

determined <strong>by</strong> gel permeation chromatography, and the covalent linkage of the cytostatic to the<br />

main polymer chain. 1 H NMR analyses demonstrated that a significant fraction of 5-FU was<br />

covalently bonded to the PBCA chains through its amino groups, preferentially through one of<br />

the two nitrogen atoms.<br />

Stella et al. 40 developed a new concept to target the folate-binding protein <strong>by</strong> designing poly<br />

(ethylene glycol) (PEG)-coated biodegradable nanoparticles coupled to folic acid. This system is<br />

the soluble form of the folate receptor that is over expressed on the surface of many tumoral cells.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


258<br />

The copolymer poly[aminopoly(ethylene glycol)cyanoacrylate-co-hexadecyl cyanoacrylate]<br />

[poly(H 2NPEGCA-co-HDCA)] was synthesized, and their nanoparticles were prepared using<br />

the nanoprecipitation technique. The nanoparticles were then conjugated to the activated folic<br />

acid via PEG terminal amino groups and purified from unreacted products. The specific<br />

interaction between the conjugate folate-nanoparticles and the folate-binding protein was<br />

confirmed <strong>by</strong> surface plasmon resonance. This interaction did not occur with the nonconjugated<br />

nanoparticles used as a control.<br />

15.2.3 CHARACTERIZATION OF NANOPARTICLES<br />

15.2.3.1 Physicochemical Characterization<br />

The methods used for the physicochemical characterization of nanoparticles are listed in<br />

Table 15.1. Particle size is the prominent feature of nanoparticles; the fastest and most routine<br />

methods of size analysis are photon correlation spectroscopy (PCS) and laser diffractometry (LD).<br />

The former method is useful for the determination of smaller particles, 41 whereas the latter method<br />

is useful for the determination of larger particles. PCS determines the hydrodynamic diameter of the<br />

nanoparticles via Brownian motion. The electron microscopy methods also allow the exact particle<br />

determination. Scanning electron microscopy requires the coating of the dry sample with a conductive<br />

material such as gold. The gold coating usually results in an estimate of particle sizes that is<br />

slightly greater than the normal. TEM, with or without staining, is a relatively easier method to<br />

determine particle size. Some nanoparticle materials are not electron dense and thus cannot be<br />

stained; others melt and sinter when irradiated <strong>by</strong> the electron beam of the microscope and thus<br />

cannot be visualized <strong>by</strong> this method. Such nanoparticles can be visualized in TEM after freeze<br />

fracture or freeze substitution. 42 This method is optimal because it also allows for the fracturing<br />

of the particles and consequently an observation of their interior. Unfortunately, the method is<br />

time-consuming and cannot be used for the routine determinations. Recently, new types of highresolution<br />

microscopes such as the atomic-force microscope, the laser-force microscope, and the<br />

scanning-tunneling microscope were developed. 43–46 These microscopes are especially useful for<br />

the investigation of nanoparticle surfaces.<br />

The molecular weight determination of nanoparticles can be performed after the particles have<br />

been dissolved in an appropriate solvent and analyzed <strong>by</strong> gel-permeation chromatography.<br />

Accurate results can be obtained only if the polymer standards have a similar structure and<br />

similar properties as the test material. The limitation of this technique, however, is that the<br />

TABLE 15.1<br />

List of Characterization Techniques for Nanoparticles<br />

Parameter Method of Characterization<br />

Nanotechnology for Cancer Therapy<br />

Particle size Photon correlation spectroscopy, laser diffraction, transmission electron microscopy,<br />

scanning electron microscopy<br />

Molecular weight Gel permeation chromatography<br />

Density Helium compression pycnometer<br />

Crystallinity X-ray diffractometry, differential scanning calorimetry<br />

Surface charge Electrophoresis, laser Doppler anemometry, amplitude-weighted phase structuration<br />

Hydrophobicity Hydrophobic interaction chromatography, contact angle measurements<br />

Surface characteristics Scanning electron microscopy, atomic force microscopy<br />

Surface element analysis X-ray photoelectron spectroscopy for chemical analysis (ESCA)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 259<br />

molecular weight of cross-linked polymers and nanoparticles from natural macromolecules cannot<br />

be determined.<br />

Density measurements can be performed <strong>by</strong> the helium compression pycnometry and <strong>by</strong><br />

density gradient centrifugation. A comparison of these two methods may offer information about<br />

the internal structure of nanoparticles.<br />

Further information about the nanoparticle structure and crystallinity may be obtained using<br />

x-ray diffraction and thermoanalytical methods such as differential scanning calorimetry, differential<br />

thermal analysis, thermogravimetry, and thermal optical analysis. 47 Thechargeon<br />

nanoparticle surfaces is mainly determined <strong>by</strong> electrophoretic mobility, laser Doppler anemometry,<br />

and amplitude-weighted phase structuration.<br />

Hydrophobicity of the nanoparticles can be determined <strong>by</strong> two major methods: contact angle<br />

measurements and hydrophobic interaction chromatography. The former can be performed only on<br />

flat surfaces and not directly on hydrated nanoparticles in their dispersion media. For this reason,<br />

hydrophobic interaction chromatography is the better method, wherein a differentiation between<br />

nanoparticles with different surface properties can be obtained <strong>by</strong> loading the particles on columns<br />

with alkyl-sepharose and eluting them with a Triton X-100 gradient. 48<br />

Kreuter 41 performed the physicochemical characterization of polymethyl, polyethyl, and<br />

PBCA nanoparticles. Because it gives only the mean diameter and the multimodal size distributions<br />

cannot be measured, determination of nanoparticle size <strong>by</strong> Photon Correlation Spectrometry is very<br />

susceptible to errors caused <strong>by</strong> the presence of bigger particles in the dispersion. Another disadvantage<br />

is that this method measures the Brownian motion of the particles, and the particle<br />

diameter is then calculated from the measurements via the diffusion coefficient. 41 Thus, the particle<br />

size determined <strong>by</strong> this method can be affected <strong>by</strong> the surrounding medium, including the adsorbed<br />

surfactants or hydration layers, and may therefore be different from the sizes obtained <strong>by</strong> electron<br />

microscopy. In case of characterization using electron microscopy, the individual particles can be<br />

analyzed and measured. According to Kopf et al., 26 the surfactants present after the separation of<br />

the dispersion medium from the nanoparticles will coat the nanoparticles and lead to the disappearance<br />

of individual nanoparticle structures. This coating significantly reduces the particle<br />

surface area, and polycyanoacrylate nanoparticles could thus not be obtained surfactant-free in a<br />

nonaggregated form. The nanoparticles were x-ray amorphous, and their diffractograms indicated<br />

no sign of crystallinity. 20,21 The surface charge of colloidal particles expressed <strong>by</strong> their electrophoreitc<br />

mobility may have an important influence on their body distribution behavior in humans<br />

and animals. 49,50 The PACA nanoparticles possess negative charge. Charge measurements in serum<br />

yielded a more pronounced decrease caused <strong>by</strong> the adsorption of the serum contents. Another<br />

parameter that is of great importance for the fate of nanoparticles in the blood is the hydrophilicity/hydrophobicity<br />

of the particle surface. The latter has a profound influence on the interaction of<br />

the particles with the blood components. 49,51,52 The particles’ hydrophilicity/hydrophobicity can be<br />

determined <strong>by</strong> measuring the water contact angle. 53 This was done <strong>by</strong> centrifugation, followed <strong>by</strong><br />

washing with water and drying the poly(cyanoacrylate) nanoparticles. The dried material was<br />

dissolved in acetone and casted as a film with subsequent solvent evaporation. 41 The water<br />

droplet contact angles for polymethyl, polyethyl, and PBCA nanoparticles were reported as<br />

55.68, 64.78, and 68.98, respectively. 41<br />

Muller et al. 48 studied the physicochemical properties of alkyl cyanoacrylates of different chain<br />

lengths such as methyl, ethyl, isobutyl, and isohexyl cyanoacrylate. Zeta potential of the nanoparticles<br />

was determined as an alternative means of evaluating particle surface charge. Interaction of<br />

nanoparticles with serum proteins was assessed <strong>by</strong> the zeta-potential measurements. The nanoparticles<br />

were incubated with poloxamer 407 (Pluronic F-108), and the hydrophobicity determination<br />

<strong>by</strong> contact angle measurement revealed that the coating layer was very thin. The layer thickness<br />

was higher in isohexyl cyanoacrylate particles, indicating that they had the most hydrophobic<br />

nature of all the polymers used.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


260<br />

The molecular weight differences between polycyanoacrylate nanoparticles with different sidechain<br />

lengths as well as the molecular weight differences of polycyanoacrylate nanoparticles<br />

produced at different hydrochloric acid concentrations or surfactant concentrations appear to not<br />

be very pronounced. 54 Nevertheless, an increase in side-chain length led to the increasing molecular<br />

weights, whereas increasing the hydrochloric acid concentrations reduced the molecular weights.<br />

15.2.3.2 Degradation of Poly(Alkyl Cyanoacrylate) Nanoparticles<br />

Nanotechnology for Cancer Therapy<br />

The type of degradation (bulk or surface degradation) of polymethyl-, polyethyl-, and poly(isohexyl<br />

cyanoacrylate) nanoparticles was determined using photon correlation spectroscopic measurements.<br />

48 In the case of surface degradation, the particle size should show an immediate increase<br />

while the polydispersity index remaining constant. If bulk degradation dominates, a lag period<br />

should occur, preceding the decrease in mean size due to disintegration of the particles. This<br />

process of disintegration would lead to a more heterogeneous distribution of particle sizes and<br />

consequently to an increase in the polydispersity index. Incubation of polyethyl cyanoacrylate<br />

nanoparticles with in 10 K4 N NaOH led to an immediate, continuous decrease in particle size,<br />

as well as an unchanged polydispersity index indicating the predominant surface degradation.<br />

Incubation of PIHCA nanoparticles in 10 K4 N NaOH led to no detectable size decrease; polydispersity<br />

was also unchanged, indicating the degradation <strong>by</strong> surface erosion at a much slower rate<br />

than poly(ethyl cyanoacrylate) nanoparticles. Coating the particles with poloxamers did not accelerate<br />

the particle degradation. At low electrolyte concentration, the size and dispersity of PIHCA<br />

nanoparticles remained unchanged during incubation in 10 K4 N NaOH. At high ionic strength, the<br />

size increase of nanoparticles as a result of flocculation was larger than the size decrease due to<br />

degradation. In vitro degradation was also determined <strong>by</strong> turbidimetric measurements. 55 Poly(methyl<br />

cyanoacrylate) and poly(ethyl cyanoacrylate) nanoparticles underwent the fastest<br />

degradation. A high electrolyte concentration in the medium led to the formation of larger aggregates<br />

accompanied <strong>by</strong> an increase in absorption of dispersion. The slowly degrading polymers<br />

PIBCA and PIHCA probably release a low concentration of degradation products over a prolonged<br />

period of time, indicating their low toxicity.<br />

The stability studies on PBCA nanoparticles reveal that an acidic medium protects the nanoparticles<br />

against decomposition. 56,57 Higher temperatures promoted the degradation of<br />

nanoparticles. When incubated in human serum in vitro, the nanoparticles showed no particle<br />

agglomeration, suggesting their better tolerance in vivo.<br />

15.2.3.3 Influence of Enzymes on the Stability of Poly(Alkyl Cyanoacrylate) Nanoparticles<br />

The peroral route is the most convenient way of delivering drugs, leading to better patient compliance,<br />

especially during long-term treatment. However, many drugs, including proteins and<br />

peptides, are unstable in the gastrointestinal tract (GIT) or are insufficiently absorbed. Earlier<br />

studies have shown that cyanoacrylate nanoparticles 58,59 and nanocapsules 60,61 improve the absorption<br />

of several drugs like vincamin or insulin from the GIT. Maincent et al. 58 demonstrated the<br />

efficiency of PBCA nanoparticles as a drug delivery system for the GIT <strong>by</strong> improving the absorption<br />

of vincamin <strong>by</strong> more than 60% compared to an equimolar solution. Investigation of the<br />

parameters that influence the stability of nanoparticles in the GIT is much more difficult because<br />

of the complexity of the GIT. The pH changes from 2.0 in the stomach to 7–8 in the small intestine.<br />

Moreover, in different parts of the GIT there are different enzymes in differing concentrations.<br />

Degradation studies with cyanoacrylate was performed as early as 1972. 62 Lenaerts et al. 63 incubated<br />

PIBCA nanoparticles in rat liver microsomes. As these microsomes contain a mixture of<br />

enzymes, it was difficult to determine which enzyme took part in the degradation process. Scherer et<br />

al. 56 studied the influence of different enzymes in vitro on the stability of PBCA nanoparticles. The<br />

study used pepsin, amylase, and esterase enzymes, added in a buffer solution that contained<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 261<br />

dispersed nanoparticles. Amylase was selected because dextran is used as a stabilizer in nanoparticle<br />

preparation and was shown to be partially incorporated into the polymer chains. 64 Esterase was<br />

selected because hydrolysis of the ester side chain is the main route of degradation of PACA<br />

nanoparticles. 65 The study showed significant degradation of PBCA nanoparticles at 378C and<br />

was dependent on the pH of the buffer solution. The addition of pepsin and amylase to the<br />

buffer solution was found to have no influence on the stability of nanoparticles. The results upon<br />

addition of amylase revealed that dextran either was not incorporated to a significant degree or that<br />

its incorporation did not alter the regular nonenzymatic or esterase-dependant degradation pathway<br />

and velocity of this polymer in the form of nanoparticles. In contrast, the esterase significantly<br />

influenced the stability of the nanoparticles. Esterase demonstrated the highest activity of nanoparticle<br />

degradation at neutral pH values; as well, the degradation of PBCA nanoparticles was<br />

proportional to the amount of esterase added. These observations suggest that the degradation of<br />

PBCA nanoparticles does not occur before the particles reach the small intestine, as esterase is<br />

released <strong>by</strong> the pancreas into the lumen of the small intestine, the major site for absorption of<br />

several drugs.<br />

15.2.3.4 Drug Release from Poly(Alkyl Cyanoacrylate) Nanoparticles<br />

Various methods were used to study the in vitro drug release of nanoparticles. They include side<strong>by</strong>-side<br />

diffusion cells with artificial or biological membranes, the dialysis-bag diffusion technique,<br />

the reverse dialysis sac technique, ultrafiltration, ultracentrifugation, and the centrifugal ultrafiltration<br />

techniques. Despite the development of these methods, some technical difficulties remain, 66<br />

including the separation of the nanoparticles from the release media when assessing the in vitro<br />

drug release from nanoparticles. Leavy and Benita 67 used a reverse dialysis method, in which the<br />

nanoparticles were added directly into the dissolution medium. The same technique was adopted <strong>by</strong><br />

Calvo et al. 68 for the evaluation of nanoparticles, nanocapsules, and nanoemulsions. However, this<br />

method is not very sensitive for studying the rapid release formulations, although relatively<br />

speaking the dialysis bag technique is the best for assessing in vitro drug release. Because the<br />

nanoparticle dispersion is placed in the dialysis bag and immersed into the release medium, this<br />

technique avoids the problem of nanoparticle separation. 69,70 Considering the hindrance to drug<br />

diffusion caused <strong>by</strong> the dialysis bag, the release of the drug in solution form can be studied through<br />

the dialysis bag, and any diffusional barrier can be compensated for.<br />

15.2.4 IN VIVO DISTRIBUTION OF POLY(ALKYL CYANOACRYLATE) NANOPARTICLES<br />

Colloidal drug carriers such as liposomes and nanoparticles are able to modify the distribution of an<br />

associated substance. They can therefore be used to improve the therapeutic index of drugs <strong>by</strong><br />

increasing their efficacy or reducing their toxicity. The polymer nanoparticles, particularly PACA<br />

nanoparticles, have attracted considerable attention as potential drug delivery systems, not only for<br />

their enhancement of therapeutic efficacy but also because they reduce the toxicity associated with a<br />

variety of drugs. Careful design of these delivery systems with respect to the target and route of<br />

administration may provide a solution to some of the delivery problems posed <strong>by</strong> many anti-cancer<br />

drugs, including doxorubicin and amphotericin B, 71 and new classes of active molecules such as<br />

peptides, proteins, genes, and oligonucleotides.<br />

The use of colloidal particulate carrier systems with controlled particle size (25 nm to 1.7 mmin<br />

diameter) is significant in such applications where diminished uptake <strong>by</strong> mononuclear phagocytes<br />

and specific targeting of carriers to particular tissues or cells is of great importance. Studies<br />

conducted on the tissue distribution of PBCA nanoparticles with a diameter of 127 nm, loaded<br />

with 1-(2-chloroethyl)-3-(1-oxyl-2,2,6,6-tetramethyl piperidinyl)-1-nitrosourea (spin-labeled nitrosourea,<br />

SLCNU) suspensions injected intraperitoneally (i.p.) into Lewis lung carcinoma-bearing<br />

mice showed a relatively low accumulation of nanoparticles in the liver and spleen; the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


262<br />

Nanotechnology for Cancer Therapy<br />

nanoparticles were mainly found in the lungs, kidneys, and heart. 72 The highest content of the<br />

particles was observed in the lungs of tumor-bearing experimental animals damaged <strong>by</strong> metastases,<br />

suggesting the possible usefulness of SLCNU–PBCA nanoparticles for targeting the<br />

lung metastases.<br />

Reddy and Murthy 73 compared the pharmacokinetics and tissue distribution of doxorubicin<br />

(Dox) solution with Dox-loaded PBCA nanoparticles synthesized <strong>by</strong> DP and EP techniques after<br />

intravenous (i.v.) and intraperitoneal (i.p.) injection in healthy rats. The elimination half-life (T1/2)<br />

and mean residence time (MRT) in blood was significantly higher than the free Dox after i.v.<br />

injection. After i.p. injection, DP nanoparticles quickly appeared in the blood and underwent rapid<br />

distribution to the organs of RES, whereas the EP nanoparticles were absorbed slowly into the<br />

blood and remained in the circulatory system for a longer time. The bioavailability (F) ofDP<br />

(w1.9-fold) and EP nanoparticles (w2.12-fold) was higher compared to that of the Dox solution<br />

after i.p. injection. The distribution to the heart of both types of nanoparticles was low after i.v. and<br />

i.p. injection, compared to Dox. This was further proved <strong>by</strong> experiments using nanoparticles<br />

radiolabeled with 99m Tc to study their distribution in Dalton’s lymphoma implanted in the thigh<br />

region of mice, when administered subcutaneously below the tumor region. 74 A significantly high<br />

tumor uptake of 99m Tc-EP nanoparticles (13-fold higher at 48 h post-injection) (P!0.001) was<br />

found compared to the 99m Tc-Dox solution.<br />

Blank [ 14 C]-poly(hexyl cyanoacrylate) nanoparticles with diameters of 200–300 nm injected<br />

intravenously into nude mice bearing a human osteosarcoma showed distribution to the liver,<br />

spleen, lung, heart, kidney, GI tract, gonads, brain, and muscle, as well as in serum and transplanted<br />

tumor fragments when investigated <strong>by</strong> liquid scintillation counting. 75 The peak levels in all organs<br />

with the exception of tumors and the spleen were reached within 24 h; the highest levels of<br />

radioactivity were found after 7 days in the tumor and spleen. The tissue distribution of naked<br />

and either normal immunoglobulin G or monoclonal antibody (antitumor osteogenic sarcoma)coated<br />

poly(hexyl-2-cyanoacrylate) nanoparticles in mice bearing human tumor xenografts showed<br />

similar results: the nanoparticles were deposited mainly in the liver and spleen, and no significant<br />

uptake was found in the tumors. 76<br />

The PACA nanoparticles exhibited rapid extravascular distribution after intravenous administration.<br />

Intravenous injection of isobutyl-2-cyanoacrylate nanoparticles resulted in high<br />

concentrations in the organs of the reticuloendothelial system such as the liver, spleen, and bone<br />

marrow. Similar observations were made <strong>by</strong> Kante et al. 77 in the case of PBCA nanoparticles<br />

injected intravenously in rats. In the case of mice with subcutaneously grafted Lewis lung carcinoma<br />

cells, the nanoparticles exhibited higher lung concentrations, indicating the use of<br />

nanoparticles in the treatment of lung metastasis. 17 However, the results were somewhat different<br />

for polybutyl-2-cyanoacrylate nanoparticles injected intravenously into rabbits. 78 The lung<br />

accumulation of nanoparticles was very low in that case, indicating that the nanoparticles<br />

(126 nm) were not mechanically filtered through the capillary bed of the lungs. The difference in<br />

lung concentrations compared to the observations of Grislain et al. 17 can be explained <strong>by</strong> the<br />

difference in particle size (254 nm). Coating with azidothymidine-loaded poly(hexyl cyanoacrylate)<br />

(PHCA) nanoparticles with polysorbate 80 led to higher brain levels of azidothymidine after<br />

intravenous injection in rats, indicating the facilitation of drug passage across the blood–brain<br />

barrier (BBB). 79 Physicochemical studies have shown that the coating of nanoparticles with<br />

block copolymers such as poloxamers and poloxamines induces a steric repulsion effect, minimizing<br />

the adhesion of particles to the surface of macrophages which in turn results in decreased<br />

phagocytic uptake and significantly higher levels in blood and non-RES organs, the brain, intestine,<br />

kidneys, etc. Borchard et al. 80 showed that coating the nanoparticles with surfactants resulted in<br />

uptake <strong>by</strong> brain endothelial cells, and that polysorbate 80 was the most effective surfactant for this<br />

purpose. Further investigations unequivocally demonstrated that polysorbate 80-coated PBCA<br />

nanoparticles loaded with enkaphalin analogue dalargin enabled the delivery of this drug across<br />

the BBB, achieving a significant pharmacological effect. Gulyaev et al. 81 reported significantly<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 263<br />

higher brain concentrations of intravenously injected polysorbate 80-coated Dox–PBCA nanoparticles<br />

in rats. Three probable mechanisms were proposed for the enhanced brain transport of<br />

polysorbate 80 coated nanoparticles: (1) endocytosis, (2) the opening of the tight endothelial<br />

junctions between the brain endothelial cells, 81 and (3) the inhibition of the P-glycoprotein<br />

efflux pump responsible for multi-drug resistance, which is a major obstacle to cancer<br />

chemotherapy. 82,83<br />

Isobutyl and isohexylcyanoacrylate nanoparticles were used <strong>by</strong> Ghanem et al. 84 as drug<br />

carriers, particularly for some anti-cancer drugs. Body distribution as well as pharmacokinetics<br />

in animals and partially in humans using a radiolabeled ( 111 In and 99m Tc) free drug and its nanoparticles<br />

showed 60–75% accumulation in the reticuloendothelial system.<br />

Skidan et al. 85 studied the antitumor efficacy of Dox loaded into polysorbate 80-coated nanoparticles<br />

to the brain. Pharmacokinetics were studied in rats after the intravenous injection of four<br />

Dox preparations: (1) solution in saline, (2) solution in polysorbate 80, intravenous (3) bound to<br />

nanoparticles, and (4) bound to nanoparticles coated with polysorbate 80. The results showed no<br />

significant difference in the body distribution between the two solution formulations, whereas the<br />

two nanoparticle formulations showed significantly decreased drug concentrations in the heart.<br />

High brain concentrations of Dox (O6 mg/g) were achieved with the nanoparticles overcoated<br />

with polysorbate 80 between 2 and 4 h, whereas the concentrations observed with the other three<br />

formulations were always negligible. Antitumor efficacy of Dox associated with polysorbate<br />

80-coated nanoparticles as assessed in rats with intracranially implanted 101/8 glioblastoma<br />

resulted in a significant 3.6-fold increase in survival, whereas administration of free Dox or<br />

blank nanoparticles did not modify the mortality when compared to the control group.<br />

Following the subcutaneous injections of metoclopramide solution (5 mg/kg) and three<br />

different metoclopramide nanosphere suspensions (10 mg/kg) on two phases in wistar rats, there<br />

was a rapid absorption, distribution, and elimination of the drug solution. The maximum drug<br />

concentration was observed after 30 min of subcutaneous injection of all the tested nanosphere<br />

formulations. 15 Polyethylcyanoacrylate-hydroxypropyl cyclodextrin showed the highest concentration,<br />

followed <strong>by</strong> polyisobutylcyanoacrylate–dextran and polyethylcyanoacrylate–dextran. The<br />

area under the curves of all nanoparticle formulations was 4.8- to 1.88-times higher than that<br />

obtained for the solution form.<br />

Earlier studies showed that polystyrene and other biodegradable microspheres delivered to the<br />

intestine are preferentially absorbed <strong>by</strong> the M-cells of the Peyer’s patches. 86,87 Intragastric administration<br />

of azidothymidine-loaded PIHCA nanoparticles resulted in high concentration in the<br />

intestinal mucosa as well as in the Peyer’s patches. 88 Once localized in the mucous or intestinal<br />

tissue, the azidothymidine could be slowly released <strong>by</strong> the enzymatic action of esterases. Very low<br />

levels of azidothymidine bound to nanoparticles were detected in the lymph, suggesting an efficient<br />

capture of particles through the Peyer’s patches resulting in an intracellular localization of the drug.<br />

As a result, some immune cells would carry azidothymidine to the mesenteric lymph nodes, the<br />

location of an important viral burden and a major replication site of HIV.<br />

PACA nanoparticles are proved to be suitable delivery systems in comparison to other colloidal<br />

systems, including liposomes. Reszka et al. 89 observed that the mitanxantrone-loaded PBCA nanoparticles<br />

showed the highest tumor concentrations in B-16 melanoma-bearing mice after intravenous<br />

injection. Four different formulations—mitaxantrone solution, mitaxantrone-loaded negatively<br />

charged liposomes (small unilamellar vesicles), 14 C-PBCA nanoparticles, and poloxamine 1508coated<br />

14 C-PBCA nanoparticles—were studied for the biodistribution and tumor accumulation in the<br />

tumor-bearing mice. The liposomal formulation showed prolonged circulation in the blood and<br />

higher accumulation in the liver and spleen after 24 h of injection, compared to the nanoparticle<br />

formulations. Despite the longer circulation in the blood, however, the liposomal formulation could<br />

not result in high concentrations in the tumors, compared to the nanoparticle formulations, which<br />

showed relatively lower circulation time in the blood, indicating the efficacy of PBCA nanoparticles<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


264<br />

in delivering mitoxantrone to tumors. Similar results involving the higher efficacy of mitoxantroneloaded<br />

PBCA nanoparticles in B-16 melanoma was reported <strong>by</strong> Beck et al. 28<br />

15.2.4.1 Surface Modification to Alter Biodistribution<br />

The potential of injectable polymeric drug carriers (nanoparticles) is compromised <strong>by</strong> their<br />

accumulation in the tissues of the mononuclear phagocytic system (MPS). Thus, the therapeutic<br />

efficacy of the drugs associated with the nanoparticles is limited to the treatment of several liver<br />

diseases. 90,91 Long-circulating nanoparticles can be obtained <strong>by</strong> their surface modification with<br />

dysopsonic polymers such as PEG. 92,93 Indeed, these hydrophilic and flexible polymers can prevent<br />

the opsonin–nanoparticle interaction, which is the first step of the recognition <strong>by</strong> the immune<br />

system. In the tumor-bearing animal models, these long-circulating nanoparticles would be able<br />

to extravasate thorough the endothelium, allowing drugs to concentrate in the tumors. 94 Recently, it<br />

was shown that the synthesis of amphiphilic PEGylated polymers allows the direct preparation of<br />

PEG-coated nanoparticles, ensuring the stability of the PEG coating layer because the PEG chains<br />

remain chemically linked to the nanoparticle core. 95–97 Otherwise, a PEG coating layer simply<br />

adsorbed onto preformed nanoparticles is desorbed in vivo. 78 Peracchia et al. 98 synthesized a novel<br />

MePEGcyanoacrylate–hexadecylcyanoacrylate (PEG–PHDCA) copolymer, and the presence of a<br />

PEG layer on the nanoparticle surface was confirmed <strong>by</strong> surface chemical analysis. 99 Poly(alkyl<br />

cyanoacrylate) nanoparticles are rapidly biodegradable and hence could be able to release the<br />

associated drug into the bloodstream in a controlled manner before liver capture could occur.<br />

For solid tumors, extravasation could be followed <strong>by</strong> a rapid degradation of the polymer, leading<br />

to a rapid drug release. 100 The cytotoxicity of PHDCA polymer was decreased after its PEGylation<br />

in the mouse macrophage cell line J774. 101 This is due to the lower extent of the particle/cell<br />

interaction, because of the steric repulsion effect of PEG chains. Biodistribution studies with the<br />

above MePEGcyanoacrylate–hexadecylcyanoacrylate copolymer nanoparticles revealed that the<br />

nanoparticles exhibited long-circulating properties. 102 No effect of the degree of PEGylation on the<br />

plasma retention was observed because increasing the copolymer from 1:5 to 1:2 did not improve<br />

the nanoparticle circulation time. The 1:5 PEG–PHDCA copolymer probably already exhibited a<br />

PEG content that was able to determine a brush PEG configuration 103 and, thus, an optimal PEG<br />

density at the particle surface for the steric repulsion to occur efficiently. Together, the nanoparticles<br />

showed low liver accumulation and high spleen uptake, representing the possibility of<br />

targeting drugs to this tissue. More recently, a rapidly biodegradable copolymer poly(methoxyethylene<br />

glycol-cyanoacrylate-co-n-hexadecylcyanoacrylate) has been developed for the preparation of<br />

stealth nanoparticles. 102,104 Similarly, Ping et al. 105 synthesized poly(methoxyethyleneglycol<br />

cyanoacrylate-co-n-hexadecyl cyanoacrylate) (PEGylated PHDCA) nanoparticles loaded with an<br />

antitumor agent salvicine. The nanoparticles showed a significant initial burst of salvicine followed<br />

<strong>by</strong> a sustained release in 7.4 pH phosphate-buffered saline, thus suggesting its usefulness in<br />

tumor therapy.<br />

15.2.5 TOXICITY OF POLYALKYL CYANOACRYLATE NANOPARTICLES<br />

Nanotechnology for Cancer Therapy<br />

Toxicity of PACA nanoparticles varies depending on the alkyl chain length in the polymer. Incubation<br />

of hepatocytes with 1% poly(methyl cyanoacrylate) nanoparticles resulted in complete<br />

perforation of cell membranes. 106,107 However, PIBCA nanoparticles showed no signs of toxicity<br />

at the same concentration. This toxicity is probably due to the presence of nanoparticle degradation<br />

products in the cytoplasm following their phagocytosis. The low toxicity of isobutyl product is due<br />

to its slower degradation than the methyl product. The toxicity and safety data on PBCA nanoparticles<br />

is somewhat contradictory. Kante et al. 106 found LD50 of 230 mg/kg of PBCA<br />

nanoparticles injected intravenously in mice. Olivier et al. 108 observed a 30–40% mortality rate<br />

in mice after the injection of 166 mg/kg of PBCA nanoparticles; some animals died even at low<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 265<br />

doses. Meanwhile, Simeonova et al., 109 in an investigation of the immunomodulating properties of<br />

PBCA nanoparticles in mice, did not report any signs of toxicity at doses as high as 200 mg/kg. A<br />

much higher LD 50 of 585 mg/kg was found for nanoparticles made of a similar polymer PHCA. 110<br />

Gelperina et al. 111 reported no mortality in rats even after administering up to 400 mg/kg of empty<br />

PBCA nanoparticles.<br />

The implantation into the tissue of monomers, which polymerize within the body in orthopedic<br />

surgery and percutaneous minimally invasive radiological procedures, was performed earlier. The<br />

polymerization was mostly exothermic, and is associated with the toxicity to the tissue. 112,113 The<br />

use of n-butylcyanoacrylate in the endovascular treatment of arteriovenous malformations has been<br />

well documented. 113,114 Implantation of n-butyl-2-cyanoacrylate (NBCA) has resulted in a good<br />

occlusive effect on the parenchyma of highly vascular tumors, with extensive intralesional necrosis.<br />

The antitumor effect of the NBCA at the implant surface on adjacent tumors appeared limited. 115<br />

Histopathological examination of a surgically removed malignant pelvic para-ganglioma after a<br />

percutaneous injection of NBCA, three days after the NBCA injection, revealed variable degrees of<br />

tumoral necrosis. 116 Areas surrounding small volumes of NBCA at the periphery of the tumor<br />

showed practically no necrosis, whereas regions inside the tumor plastified with NBCA presented<br />

greater degrees of necrosis, especially toward the center of the tumoral mass (Figure 15.3).<br />

However, no necrosis was found at a macroscopic level outside the implant, indicating that the<br />

NBCA allowed for the destruction of tumor tissue within the heavily plastified areas. The carcinogenic<br />

potential of cyanoacrylate adhesive and isobutyl-2-cyanoacrylate were also reported<br />

respectively <strong>by</strong> Matsumoto 117 and Samson and Marshal. 118 Induction of malignant fibrous histiocytoma<br />

in female Fisher rats <strong>by</strong> implantation of foreign bodies was studied <strong>by</strong> Hatanaka et al. 119<br />

Five kinds of foreign bodies (silicone, cellulose, polyvinyl chloride, and zirconia and alkyl-alpha<br />

cyanoacrylate) were implanted into the subcutaneous tissue of female Fisher rats. The tumors<br />

developed in almost all the cases was composed of a mixture of cells that resembled fibroblast,<br />

Brain concentration (μg/g)<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

–60<br />

0<br />

60 120 180 240 300 360 420 480<br />

Time [min]<br />

FIGURE 15.3 Uptake <strong>by</strong> rat brain of i.v. injected doxorubicin loaded to polysorbate 80-coated nanoparticles.<br />

Doxorubicin levels in rat brain (mg/g) versus time (min). C Doxorubicin (5 mg/kg) in saline. $ Doxorubicin<br />

(5 mg/kg) bound to nanoparticles overcoated with 1% polysorbate 80. (From Gulyaev, A. E. et al., Pharm.<br />

Res., 16, 1564, 1999. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


266<br />

Golgi apparatus, histiocytes, myofibroblasts, and immature mesenchymal cells. A rat malignant<br />

fibrous histiocytoma transplanted into the subcutaneous tissue of the syngeneic female Fisher rats<br />

grew and metastasized to the lungs. Canter et al. 120 reported the consecutive pathological findings<br />

of a patient who underwent surgery for facial hemangioma after percutaneous injection of NBCA<br />

for devascularization of a lesion. The patient underwent additional surgery one month and six<br />

months after the initial operation for the removal of the residual NBCA cast from the injection<br />

site. Acute inflammatory findings after injection of NBCA and the development of a chronic<br />

granulomatous foreign-body reaction support the histological findings of experimental animal<br />

studies and postmortem examinations of humans. Though the alkylcyanoacrylates of smaller<br />

alkyl chain lengths exhibited considerable toxicity, the alkylcyanoacrylates of larger chain<br />

lengths such as isohexylcyanoacrylate nanoparticles resulted in negligible toxicity 121 and drove<br />

them to their clinical trials.<br />

15.2.5.1 Toxicity of Drugs Associated with Poly(Alkyl Cyanoacrylate) Nanoparticles<br />

The study of the acute toxicity of doxorubicin associated with polysorbate 80-coated nanoparticles<br />

in healthy rats and rats with intracranially implanted 101/8 glioblastoma was reported <strong>by</strong> Gelperina<br />

et al. 111 Single intravenous administration of empty PBCA nanoparticles in the dose range of 100–<br />

400 mg/kg did not cause mortality within the period of observation and also did not affect body<br />

weight or the weight of the internal organs. Association of doxorubicin with PBCA nanoparticles<br />

did not produce significant changes of quantitative parameters of acute toxicity of the antitumor<br />

agent. Efficacy and toxicity of mitoxantrone-loaded PACA nanoparticles were compared with a<br />

drug solution and with a mitoxantrone-liposome formulation. 89 Furthermore, the influence of an<br />

additional coating surfactant, poloxamine 1508, which has been shown to change the body distribution<br />

of other polymeric nanoparticles, was investigated. PACA nanoparticles led to a significant<br />

reduction in tumor volume of B-16 melanoma in mice compared to liposomes. However, neither<br />

nanoparticles nor liposomes had overcome the mitoxantrone-associated leucocytopenia.<br />

In a study <strong>by</strong> Gibaud et al. 122 involving the determination of the myelosuppressive effects of<br />

free and PACA-bound doxorubicin in mice in vivo, the alkyl cyanoacrylate nanoparticle-bound<br />

doxorubicin showed the highest myelosuppressive effect. The total and differential counts of blood,<br />

bone marrow, and spleen cells, as well as the number of granulocyte progenitors (CFU-GM), was<br />

determined <strong>by</strong> culture after the intravenous injection of 11 mg/kg body weight of doxorubicin,<br />

either free or bound to PIBCA or PIHCA nanoparticles. Doxorubcin bound to PIHCA nanoparticles<br />

showed the highest and longest myelosuppressive effects, which correlated well with a high<br />

concentration of the drug in the bone marrow and spleen. Furthermore, the PIHCA nanoparticles<br />

induced the release of colony-stimulating factors, which might account for the observed increase in<br />

the toxic effects of doxorubicin on bone marrow progenitors.<br />

15.2.6 DRUG DELIVERY IN CANCER WITH POLY(ALKYL CYANOACRYLATE) NANOPARTICLES<br />

15.2.6.1 Targeting to Cancer Cells and Tissues<br />

Nanotechnology for Cancer Therapy<br />

Increase in the antitumoral activity of many anti-cancer drugs was also observed when loaded in<br />

polyisohexylcyanoacrylate nanoparticles. A study 123 conducted with doxorubicin-loaded polyisohexylcyanoacrylate<br />

nanoparticles administered i.v. to the C57BL/6 mice consisting of metastases<br />

induced <strong>by</strong> i.v. inoculation of reticulosarcoma M5076 cell suspension showed a dramatic increase<br />

in the antitumoral activity of the drug. Doxorubicin measurements in healthy hepatic or neoplastic<br />

tissue, carried out with histological examinations using TEM, demonstrated that the hepatic tissue<br />

is an efficient reservoir of the drug when it was injected in association with nanoparticles. Accumulation<br />

of doxorubicin-associated nanoparticles suggests that the Kupffer cells created a gradient of<br />

drug concentration for a massive and prolonged diffusion of the free drug toward the neoplastic<br />

tissue. The negatively charged mitoxantrone PBCA nanoparticles (DHAQ-PBCA-NP) prepared <strong>by</strong><br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 267<br />

EP method 124 following intravenous administration were concentrated mainly in liver tumors. The<br />

accumulation of DHAQ-PBCA-NP was higher in the liver tumors than in the liver tissue. Yang et<br />

al. 125 compared antitumor activity of mitoxantrone (DHAQ) and mitoxantrone-polybutyl cyanoacrylate-nanosphere<br />

(DHAQ-PBCA-NS) against experimental liver tumor H22 in mice. The results<br />

of antitumor activity determination of both drugs with different treatment schedules showed<br />

DHAQ-PBCA-NS to present higher activity than DHAQ; DHAQ-PBCA-NS is possessed of<br />

liver-targeting property.<br />

Soma et al. 126 demonstrated that the Doxorubicin-loaded PACA nanoparticles are more<br />

efficient than free drugs in mice bearing hepatic metastasis of the M5076 tumor. High phagocytic<br />

activity of Kupffer cells in the liver played the role of drug reservoir after nanoparticle phagocytosis.<br />

The study also assessed the role of macrophages in mediating the cytotoxicity of doxorubicin-loaded<br />

nanoparticles on M5076 cells. After the phagocytosis of the doxorubicin-loaded nanoparticles,<br />

J774.A1 cells were able to release the active drug, allowing it to exert its cytotoxicity against<br />

M5076 cells. Intracellular accumulation and DNA binding of doxorubicin encapsulated in PIHCA<br />

nanospheres and of daunorubicin bound to polyglutamic acid (DGA) in comparison with free Dox<br />

and daunorubicin was studied <strong>by</strong> Bogush et al. 127 using methods involving interaction between an<br />

anthracycline and cellular DNA to understand how the drug reaches its nuclear targets. The results<br />

showed that the intracellular accumulation and DNA binding of Dox-loaded nanospheres and DGA<br />

are reduced <strong>by</strong> 30–40% in comparison with those obtained for free doxorubicin and<br />

daunorubicin, respectively.<br />

Studies <strong>by</strong> Bennis et al. 128 on the cytotoxicity and accumulation of doxorubicin encapsulated in<br />

PIHCA nanospheres in a model of doxorubicin-resistant rat glioblastoma variants revealed that the<br />

cytotoxicity differed with the degree of resistance to this drug. The nanoparticle associated Dox was<br />

more cytotoxic than the free Dox, whereas co-administration of drug-free nanoparticles with free<br />

Dox did not significantly modify the cytotoxicity of Dox. However, Simeonova and Antcheva 129<br />

reported that the cytotoxic activities of drug-free PBCA nanoparticles free of vinblastine, vinblastine-loaded<br />

nanoparticles (<strong>by</strong> incorporation and adsorption processes), and a mixture of vinblastinefree<br />

and drug-free PBCA nanoparticles, when compared in vitro on human erythroleukemic K-562<br />

cells, showed enhanced cytotoxicity when vinblastine was either adsorbed on PBCA or mixed with<br />

them rather than free. When vinblastine was incorporated into the polymer matrix of nanoparticles,<br />

a lag period and a postponed cytotoxic effect on K-562 cells was observed. Preclinical studies 121 in<br />

21 patients with refractory solid tumors also confirmed an increase in doxorubicin cytotoxicity and<br />

a decrease in cardiotoxicity when incorporated into biodegradable PIHCA <strong>by</strong> modifying tissue<br />

distribution. In clinical studies 130 conducted on 21 patients with Stages III–IV maxillary tumors,<br />

antitumor drugs deposited with a cyanoacrylate adhesive solution showed immediate and late<br />

results of therapy. This evidence of deposition of the drug and its regulated elimination from the<br />

composition helps create a high concentration of the agent at the operation site where the tumor<br />

elements remained after surgery.<br />

Pan et al. 131 demonstrated the higher efficacy of 5-FU when loaded into PBCA nanocapsules<br />

with a diameter of 338 nm, compared to the free drug in S180, HepS, and EAC tumors.<br />

The in vitro and in vivo antihepatoma effects of the liver-targeted drug delivery system lyophilized<br />

aclacinomycin-A loaded polyisobutylcyanoacrylate nanoparticle (ACM-IBC-NP) was found<br />

to be significantly concentration dependent. 132 The antihepatoma activity of lyophilized<br />

ACM-IBC-NP was higher than that of the aclacinomycin-A. Similarly, actinomycin-D absorbed<br />

into polymethylcyanoacrylate nanoparticles showed increased efficiency against an experimental<br />

tumor in comparison to the free drug. 133<br />

The antitumor efficacy of a medical-grade cyanoacrylate adhesive MK-8-based composition<br />

with prospidin has been studied <strong>by</strong> Chissov et al., 134 and the kinetics of the drug elimination have<br />

been followed up in a rat model. The findings showed a prolongation of the relapse-free period and<br />

a longer survival of the experimental animals. Clinical trials of the composition carried out in 70<br />

patients with esophageal carcinomas showed local disseminated forms in 52 patients. Application<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


268<br />

of the adhesive composition has not increased the rate of postoperative complications; in some<br />

cases, the life span of patients after palliative surgery was considerably longer. In a study <strong>by</strong> Pérez<br />

et al. 135 using an N-butyl-2-cyano acrylate-based tissue adhesive, Tisuacryl, as a nonsuture method<br />

for closing wounds in oral surgery involving apicectomy, molar extractions, and mucogingival<br />

grafting, the adhesive was well tolerated <strong>by</strong> the tissue and permitted immediate hemostasis and<br />

normal healing of incisions.<br />

Association of 5-FU to PBCA nanoparticles also showed enhanced efficacy against Crocker<br />

sarcoma S 180 and a higher toxicity of the drug, as measured <strong>by</strong> induced leukopenia, body weight<br />

loss, and premature death. 31 The efficacy was further increased <strong>by</strong> an increase in the polymerto-drug<br />

ratio. The nanoparticles yielded a prolonged persistence of the 5-FU in all organs examined,<br />

including the tumor.<br />

15.2.6.2 Drug Delivery to Resistant Tumor Cells<br />

Nanotechnology for Cancer Therapy<br />

Several cancers show resistance to conventional chemotherapy. Resistance to cancer chemotherapy<br />

involves both altered drug activity at the designated target and modified intra-tumor pharmacokinetics<br />

(e.g., uptake and metabolism). The membrane transporter P-glycoprotein plays a major role<br />

in pharmacokinetic resistance <strong>by</strong> preventing sufficient intracellular accumulation of several anticancer<br />

agents. Inhibiting the P-glycoprotein has great potential to restore chemotherapeutic effectiveness.<br />

136 PACA nanoparticles were found to overcome multi-drug resistance (MDR) phenomena<br />

at both the cellular and noncellular level. 137,138<br />

Doxorubicin is a well known P-glycoprotein substrate. The resistant cells treated with doxorubicin-loaded<br />

PACA nanoparticles showed much higher sensitivity to the drug relative to the free<br />

doxorubicin. 139 Dox and Dox-loaded PIHCA nanoparticles were tested in two human tumor cell<br />

lines, K562 and MCF7, at increasing concentrations. The cell lines were more resistant to the free<br />

Dox than to the Dox-loaded nanoparticles. However, the MCF7 sublines showed a higher level of<br />

resistance to Dox-loaded nanoparticles than that of the free Dox. Different levels of overexpression<br />

of several genes involved in drug resistance (MDR1, MRP1, BCRP, and TOP2alpha) occurred in<br />

the resistant variants. MDR1 gene overexpression was consistently higher in free-Dox selected cells<br />

than in Dox-loaded nanoparticle-selected cells; this was the reverse for the BCRP gene. Overexpression<br />

of the MRP1 and TOP2alpha genes was also observed in the selected variants. The<br />

results indicate that several mechanisms may be involved in the acquisition of drug resistance and<br />

that drug encapsulation markedly alters or delays these processes. 140<br />

Further experiments, such as uptake studies in the presence of cytochalasin B or efflux studies,<br />

indicated a possible mechanism of nanoparticle–cell interaction. The degradation of the drug carrier<br />

was also shown to play a key role in the mechanism of action. The formation of an ion pair <strong>by</strong> the<br />

polycyanoacrylic acid resulting from the nanoparticle degradation with doxorubicin 141 has been<br />

considered as one of the possibilities (Figure 15.4). It was hypothesized that the nanoparticles<br />

adhere to the cell surface, followed <strong>by</strong> the release of doxorubicin and nanoparticle degradation<br />

products, which combine as an ion pair able to cross the cell membrane without being recognized<br />

<strong>by</strong> the P-glycoprotein (Figure 15.5). 142,143 The ion pair was considered to form between the<br />

positively charged doxorubicin and the negatively charged polycyanoacrylic acid, resulting in<br />

the charge masking of the doxorubicin and there<strong>by</strong> making it more lipophilic and able to cross<br />

the cell membrane. The formation of an ion pair of 2-cyano-2-butyl hexanoic acid (a PACA-like<br />

molecule) with the doxorubicin has been assessed using ion pair reverse-phase high-performance<br />

liquid chromatography. As the PACA possess high molecular weight (w1000) and their nanoparticles<br />

possess large polydispersity, a PACA-like molecule, i.e., 2-cyano-2-butyl hexanoic acid<br />

(molecular weight of 197.28 g/mol) was used to investigate the formation of the ion pair. The<br />

ion pair formation was found between the charged amino group of doxorubicin and the carboxyl<br />

group of 2-cyano-2-butyl hexanoic acid, leading to the overall decrease in charge of the drug and<br />

the increase in drug lipophilicity (Figure 15.6). The experimental results were justified based on two<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 269<br />

FIGURE 15.4 Histological section of resected paraganglioma after percutaneous NBCA injection. (a) Section<br />

close to the tumor surface showing NBCA (void) and viable tumor around the implant (H&E,<br />

original magnification X25). (b) High-power view of (a). Note absence of necrosis around the implant<br />

(H&E, original magnification X100). (c) Section in a deeper zone of the tumor. Note extensive necrosis<br />

(H&E, original magnification X25). (d) High-power view of (c). Note an acellular necrotic area (H&E, original<br />

magnification X100). (From San Millan Ruiz, D. et al., Bone, 25(Suppl. 2), 85S, 1999. With permission.)<br />

assumptions: (1) the 2-cyano-2-butyl hexanoic acid and PACAs behave similarly in terms of acidic<br />

properties, and (2) the organic and aqueous buffer mixtures used in the chromatographic procedure<br />

could be correlated with the physiological medium.<br />

In another experiment, 142 C6 rat glioblastoma cells—the sensitive, resistant C6 0.001 (continuously<br />

grown in 0.001 mg doxorubicin/mL medium) and the C6 0.5 cells (continuously grown in<br />

0.5 mg doxorubicin/ml medium)—and the drug/formulation were separated <strong>by</strong> a polycarbonate<br />

membrane (pore size 0.2 mm) (Figure 15.7). The C6 sensitive cells showed the same cytotoxicity<br />

in the presence or absence of the membrane. While the doxorubicin-loaded PIHCA nanoparticles<br />

were found to overcome the resistance of C6 0.001 and C6 0.5 compared to the free drug and free<br />

drug C blank nanoparticles when in direct contact with the drug, a 4-fold and 5-fold resistance,<br />

respectively, was maintained when separated <strong>by</strong> the membrane. This clearly indicates the necessity<br />

of contact of the drug-loaded nanoparticles with the cell membranes to overcome the MDR in<br />

resistant cells. To confirm the possible interference of the nanoparticle degradation products with<br />

P-glycoprotein, the inhibition study of tritiated azidopine binding to the P-glycoprotein enriched<br />

membranes of C6 0.5 cells was performed. The blank nanoparticles could not inhibit the azidopine<br />

binding even after spontaneous degradation, whereas both the free doxorubicin and its nanoparticle<br />

formulation decreased the binding, indicating that the degradation products of nanoparticles do not<br />

have a direct role in P-glycoprotein (P-gp) inhibition and that it is the combination of both the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


270<br />

nucleus<br />

nanoparticle<br />

P-gp<br />

(a) (b)<br />

Nanoparticle in the<br />

course of degradation<br />

nucleus<br />

Ion pair<br />

Doxorubicin<br />

FIGURE 15.5 Hypothesis about the mechanism of action of PACA nanoparticles to overcome MDR at the<br />

cellular level. Drug-loaded nanoparticles are not endocytosed <strong>by</strong> the resistant cells (a) but adhere to the cell<br />

surface where they degrade and simultaneously release degradation products and the drug (b). The degradation<br />

products and the drug form ion pairs (c) that can penetrate the cells without being recognized <strong>by</strong> the P-gp and,<br />

<strong>by</strong> this means, increase the intracellular concentration of the anti-cancer drug in the resistant cells. (From<br />

Vauthier, C. et al., J. Control. Release, 93, 151, 2003. With permission.)<br />

doxorubicin and the PACA in the nanoparticle formulation that plays a role in<br />

P-glycoprotein inhibition.<br />

Among the various strategies implemented for <strong>by</strong>passing the MDR in cancer cells (such as the<br />

use of ribozymes, 144 oligonucleotides, 145 and chemosensitizing compounds), the co-administration<br />

of chemosensitizing compounds with the anti-cancer agents has been widely investigated. 146<br />

Of such compounds, cyclosporine A, a potent immunosuppressive agent, was shown to prevent<br />

MDR in resistant cells in culture 147 and exhibited promising clinical results in refractory<br />

leukemia. 148 Cyclosporine A is believed to bind directly to P-gp, there<strong>by</strong> inhibiting the efflux<br />

pump of the cytotoxic agents, resulting in their higher intracellular accumulation. Co-encapsulation<br />

of cyclosporine A and doxorubicin into polyisobutyl cyanoacrylate nanoparticles has resulted in<br />

a significant improvement in the reversion of MDR in P388-resistant cells compared to<br />

the mixed solution of doxorubicin and cyclosporine A and the nanoparticle-loaded doxorubicin<br />

incubated with free cyclosporine A. 149 The nanoparticles target the loaded drugs close to the cell<br />

membrane, providing the higher drug concentrations at the cell surface, 150 and this could be the<br />

reason for the higher efficacy of drugs loaded into the nanoparticles compared to those in<br />

solution form.<br />

Co-administration of cyclosporine A with anti-cancer agents was shown to overcome the multidrug<br />

resistance in resistant cell lines. 147 To understand the contribution of macrophages in delivering<br />

the doxorubicin-loaded nanoparticles to the tumor cells, a co-culture of sensitive or resistant<br />

P388 tumor cells and macrophage monocyte cells J774.A1 were incubated with the free doxorubicin,<br />

doxorubicin-loaded isobutylcyanoacrylate nanoparticles, and a combined cyclosporine A<br />

and doxorubicin-loaded nanoparticle formulation. 151 When incubated directly with tumor cells,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

+<br />

(c)


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 271<br />

(a)<br />

+ +<br />

− −−<br />

Doxorubicin<br />

O<br />

OCH3 O<br />

H 3 C<br />

+<br />

OH<br />

OH H O<br />

OH<br />

O<br />

NH 3+<br />

+ + +<br />

− −−<br />

+ +<br />

O<br />

C CH2OH OH<br />

O<br />

−−−<br />

CN<br />

−<br />

+ +<br />

−−<br />

HO<br />

CH 3 CH 3 ONa<br />

Butylbromide<br />

OCH2CH3 (1)<br />

O<br />

OH<br />

(1) Ethyl cyanoacetate<br />

(2) Ethyl (2-cyano 2-butyl) hexanoate<br />

(3) (2-cyano 2-butyl) hexanoic acid<br />

(b)<br />

CN<br />

+<br />

−−−<br />

PACA degradation product<br />

−−−<br />

CH 2<br />

C 4 H 9<br />

C 4 H 9<br />

(3)<br />

O<br />

CN<br />

C<br />

C<br />

O −<br />

H<br />

O<br />

n=3<br />

CN<br />

C 4 H 9<br />

C 4 H 9<br />

OCH 2 CH 3 (2)<br />

FIGURE 15.6 2-Cyano-2-butyl hexanoic acid and doxorubicin. (a) Hypothesis on the biological action of<br />

PACA ion pair formation with doxorubicin. (b) Structure of the hypothetic ion pair formation and synthesis of<br />

the 2-cyano-2-butyl hexanoic acid (From Pepin, X. et al., J. Chromatogr. B, 702, 181, 1997. With permission.)<br />

medium only<br />

Cell monolayer<br />

KOH<br />

drug + medium Insert<br />

Well of a 24-well<br />

microplate<br />

polycarbonate membrane, 0.20 mm<br />

FIGURE 15.7 Schematic representation of the experimental device used for separating cell and drug compartments<br />

<strong>by</strong> a polycarbonate membrane (pore size 0.20 km). (From Hu, Y.-P. et al., Cancer Chemother.<br />

Pharmacol., 37, 556, 1996. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


272<br />

both the doxorubicin and the doxorubicin-loaded nanoparticle formulation were equally cytotoxic.<br />

However, in co-culture, the doxorubicin-loaded nanoparticle formulation was more effective than<br />

the free doxorubicin. In co-culture, whereas the nanoparticle-loaded doxorubicin enhanced the<br />

cytotoxicity compared to the free drug, it has partly overcome the multidrug resistance in resistant<br />

cells. The addition of cyclosporine A improved the cytotoxicity of doxorubicin irrespective of the<br />

formulation, but it could not overcome the MDR. However, the efficacy of cyclosporine A in<br />

overcoming the MDR was higher in the case of the combined nanoparticle formulation with the<br />

doxorubicin, compared to the effect seen with the mixed nanoparticles of cyclosporine A and the<br />

doxorubicin prepared separately. Such a combinatorial nanoparticulate system would have both<br />

enhanced efficacy against the tumor and be able to overcome the adverse effects associated with the<br />

administration in the free form.<br />

Doxorubicin-14-O-hemiadipate (a 14-O-hemiadipate of doxorubicin) (H-Dox) was shown to<br />

partially circumvent the P-gp in MCF7/R human breast carcinoma and P388/R murine leukemia<br />

cell lines. 152 The comparative studies of the accumulation, retention, and intracellular localization<br />

of H-Dox and Dox in Dox-sensitive murine leukemia cell line P388/S and its resistant variant, Pgppositive<br />

drug-resistant P388/R subline, in the presence or absence of cyclosporine A, showed a<br />

38-fold decrease in Dox accumulation but only a 5-fold decrease in H-Dox accumulation, indicating<br />

a greater than 7-fold increase in H-Dox buildup in resistant cells. Coincubation with<br />

cyclosporine A resulted in a 54-fold increase in Dox accumulation and only a 5-fold increase in<br />

H-Dox uptake in P388/R cells, restoring the doxorubicin levels in P388/R to 100% of that found in<br />

P388/S cells. Once internalized <strong>by</strong> the resistant cells, H-Dox was retained better than Dox, regardless<br />

of the presence or absence of CsA. The H-Dox was localized in the nuclei of the P388/R cells<br />

but not the Dox. 153 Coadministration of chemosensitizers like the cyclosporine analogue SDZ PSC<br />

833 [(3 0 -keto-Bmt1)-(Val2)-cyclosporin] (PSC 833) and anti-cancer drugs has been shown to<br />

possess powerful chemosensitization properties in vitro, in addition to being intrinsically<br />

nontoxic. 154<br />

Complete reversion of drug resistance in vitro was demonstrated <strong>by</strong> Cuvier et al. 137 with<br />

Doxorubicin-loaded PACA nanoparticles in terms of cell growth inhibition comparable with that<br />

obtained with sensitive cells exposed to free Dox. In vivo experiments showed significantly<br />

prolonged survival of P388-Adr-resistant leukemia bearing mice that had previously received<br />

P388-Adr-R cells <strong>by</strong> i.p. injections of Dox-NS, whereas free Dox injection was ineffective<br />

against this rapidly growing tumor. Intracellular concentration and the cytoplasmic and nuclear<br />

distribution of Dox as measured <strong>by</strong> laser microspectrofluorometry (LMSF) showed a similar<br />

accumulation and distribution of Dox, regardless of the form of Dox delivery in the sensitive<br />

cells, whereas Dox-NS accumulated the same amount in resistant cells.<br />

Use of folate receptor-specific drug conjugates and their nanoparticles/liposomes forms is<br />

another strategy for <strong>by</strong>passing multi-drug resistant efflux pumps. Folic acid, attached to polyethyleneglycol-derivatized,<br />

distearoyl-phosphatidylethanolamine, was used <strong>by</strong> Goren et al. 155 to target<br />

liposomes to folate receptor (FR)-overexpressing tumor cells in vitro. Additional experiments with<br />

DOX-loaded, folate-targeted liposomes (FTLs) indicate that liposomal DOX is rapidly internalized,<br />

released in the cytoplasmic compartment, and, shortly thereafter, detected in the nucleus, the entire<br />

process lasting 1–2 h. FR-mediated cell uptake of targeted liposomal DOX into a multi-drug<br />

resistant subline of M109-HiFR cells (M109R-HiFR) was unaffected <strong>by</strong> P-glycoprotein-mediated<br />

drug efflux, in sharp contrast to the uptake of free DOX.<br />

15.2.6.3 Drug Delivery to Hepatocellular Carcinomas<br />

Nanotechnology for Cancer Therapy<br />

Hepatocellular carcinomas (HCCs) are generally identified clinically at an advanced stage,<br />

usually in combination with cirrhosis. Though optimal, surgical resection is associated with<br />

a high rate of recurrence. Approaches to prevent recurrence, like chemoembolization after<br />

surgery, have not proven to be beneficial. Both doxorubicin and cisplatin are frequently<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 273<br />

used in the treatment of these carcinomas, but the overall response rates are low and neither<br />

approach seems to prolong the survival. Newly emerging agents with promising results include<br />

90<br />

Y microspheres, antiangiogenesis agents, inhibitors of growth factors and their receptors, and<br />

K vitamins. 156<br />

HCC is known to be chemoresistant to anti-cancer drugs due to the expression of multi-drug<br />

resistant (MDR) transporters. Doxorubicin loaded into the PACA nanoparticles enhanced the<br />

efficiency of treatment in the M5076 murine hepatic tumor metastasis in mice. 91 After intravenous<br />

injection, the doxorubicin-loaded PBCA nanoparticles resulted in higher hepatic concentrations in<br />

tumor-bearing mice. Histology of the mouse liver after the intravenous injection of the nanoparticle<br />

formulation revealed greater accumulation of nanoparticles in the Kupfer cells but not in the tumor<br />

cells. 123 This was supposed to be the result of macrophage uptake and subsequent capture <strong>by</strong> the<br />

Kupfer cells, of the nanoparticle formulation owing to the hydrophobicity of the nanoparticles, and<br />

then the drug release close to the tumor cells in the liver. To better understand the role of macrophages<br />

in delivering the nanoparticle-bound doxorubicin to the tumor cells, a co-culture system was<br />

developed consisting of two compartments separated <strong>by</strong> a porous polyester membrane (Transwell<br />

clear insert). 126 The results were compared with the control experiments involving direct incubation<br />

of drug/formulations with the tumor cells. The M5076 cells were seeded into the lower<br />

compartment, whereas the J774.A1 macrophage monocyte cells were placed on the top of the<br />

membrane. The doxorubicin and doxorubicin-loaded isobutylcyanoacrylate nanoparticle<br />

formulation were added to the upper compartment containing macrophages. The doxorubicin<br />

and nanoparticle formulation showed a 5-fold increase in IC50 of M5076 tumor cells in the<br />

co-culture system. In some groups of experiments, the recombinant mouse interferon-g (IFNg)<br />

was added to the macrophage compartment before the addition of the formulation to activate<br />

the macrophages. This would result in the release of cytotoxic factors such as IFN-a and nitric<br />

oxide. 157 Doxorubicin and nanoparticle formulation were found to be more cytotoxic when the<br />

macrophages were activated <strong>by</strong> IFN-g. In fact, this would probably be a result of a synergistic<br />

effect of the cytotoxic factors, especially nitric oxide released <strong>by</strong> the activated macrophages and<br />

the nanoparticle formulation because the role of antitumor factors in the inhibition of tumors both<br />

in vitro and in vivo has been well documented. 158,159 This co-culture system gives a clear<br />

indication of the role that macrophages play after intravenous injection in enhancing the<br />

cytotoxicity of the doxorubicin-loaded PACA nanoparticle formulation and its efficacy in<br />

liver tumor.<br />

Recently, Barraud et al. 160 compared the antitumor efficacy of Dox-loaded PIHCA nanoparticles<br />

with free doxorubicin, both in vitro and in vivo, in HCC-bearing transgenic mice<br />

overexpressing the mdr1 and mdr3 genes. The 50% inhibition concentration (IC50) of doxorubicin<br />

in vitro on different human hepatoma cell lines decreased with the nanoparticle formulation,<br />

compared to that of free Dox. Dox-loaded nanoparticles showed an enhanced MDR reversal in<br />

these cells.<br />

Mitoxantrone-loaded PBCA nanoparticles showed higher tumor concentration after intravenous<br />

injection in the B-16 melanoma-bearing mice. 89 These nanoparticles have also shown higher<br />

efficacy in the treatment of B-16 melanoma. 28 The efficacy of mitoxantrone-loaded PBCA nanoparticles<br />

after intravenous administration was also tested in heterotopic and orthotopically<br />

transplanted human HCC-bearing mice. 161 The efficacy of the mitoxantrone-loaded nanoparticles<br />

was compared with that of free mitoxantrone and doxorubicin. In orthotopically transplanted<br />

mice, all three preparations produced a similar effect. Microscopic examination revealed that the<br />

mitoxantrone-loaded nanoparticles showed enhanced antitumor activity in heterotopically transplanted<br />

mice, which was supported <strong>by</strong> a low tumor cell proliferation. The nanoparticles also did not<br />

show any signs of toxicity of the heart, spleen, lung, and kidney, indicating their effectiveness<br />

in tumor therapy and that the fact that they had overcome the adverse effects associated with the<br />

anti-cancer agents.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


274<br />

15.2.6.4 Drug Delivery to Brain Tumors<br />

Nanotechnology for Cancer Therapy<br />

Brain tumors are one of the most lethal forms of cancer. Many primary brain tumors are most<br />

resistant to chemotherapy, probably due to the presence of a tight BBB. 162 They are extremely<br />

difficult to treat due to a lack of therapeutic strategies capable of overcoming barriers for effective<br />

delivery of drugs to the brain. In fact, the vasculature of gliomas possess some special features:<br />

open endothelial gaps (inter-endothelial junctions and transendothelial channels having diameter of<br />

about 0.3 mm), fenestrations (5.5 nm maximum width), cytoplasmic vesicles such as caveolae (50–<br />

70 nm diameter), and vesicular vacuolar organelles (108G32 nm diameter). All of these features<br />

are due to the secretion of a vascular endothelial growth factor (VEGF) that causes the loss of the<br />

barrier function of BBB, which is essential for delivering drugs to the brain. 163 Nanoparticles are<br />

considered suitable for delivering drugs to the brain, but they should possess the following ideal<br />

characteristics: 164 (1) a nontoxic, biodegradable, and biocompatible nature; (2) a particle diameter<br />

less than 100 nm; (3) exhibit physical stability in the blood (no aggregation); (4) avoidance of the<br />

MPS (no opsonization) prolonged blood circulation time; (5) a BBB-targeted brain delivery<br />

(receptor-mediated transcytosis across brain capillary endothelial cells); (6) a scalable and costeffective<br />

manufacturing process; (7) be amenable to small molecules, peptides, proteins, or nucleic<br />

acids; (8) exhibit minimal nanoparticle excipient-induced drug alteration (chemical degradation/<br />

alteration, protein denaturation); and (9) show possible modulation of drug release profiles.<br />

Extensive efforts have been made to develop novel strategies to overcome the obstacles for<br />

brain tumor drug delivery. 165 PACA nanoparticles enabled the delivery of a number of drugs,<br />

including doxorubicin, loperamide, tubocurarine, the NMDA receptor antagonist MRZ 2/576,<br />

and the peptides dalargin and kytorphin across the BBB after coating with surfactants. 166<br />

However, only the surfactants polysorbate (Tween) 20, 40, 60, and 80, as well as some poloxamers<br />

(Pluronic F68), can induce this uptake. The mechanism for the delivery across the BBB most likely<br />

is endocytosis via the LDL receptor <strong>by</strong> the endothelial cells lining the brain blood capillaries after<br />

the injection of nanoparticles into the blood stream. This endocytotic uptake seems to be mediated<br />

<strong>by</strong> the adsorption of apolipoprotein B and/or E from the blood. After this process, the nanoparticleassociated<br />

drug may be released into the cells that diffuse into the internal parts of brain or the<br />

particles may be transcytosed.<br />

Unfortunately, the conventional approaches resulted in only sub-therapeutic concentrations in<br />

the brain, due to the short plasma half-life of the compounds as well as the carrier system. Surface<br />

coating the carrier system with polyethylene glycol led to the enhancement of plasma residence<br />

time, due to the formation of a steric barrier preventing the interaction of the carrier with the plasma<br />

proteins (opsonins) and thus preventing the opsonization and further uptake <strong>by</strong> the MPS. 92,101<br />

PEG-coated-polyalkylcyanoacrylate nanoparticles consisting of the amphiphilic copolymer<br />

poly(PEGCA-co-hexadecyl cyanoacrylate) showed longer blood circulation time and better brain<br />

penetration than the nanoparticles coated with polysorbate 80 after intravenous injection in mice<br />

and rats. Unlike the polysorbate 80-coated nanoparticles, the penetration of poly(PEGCA-cohexadecyl<br />

cyanoacrylate) nanoparticles into the brain occurred without any modification of the<br />

BBB. The brain permeability of the nanoparticles further increased in pathological conditions,<br />

where the permeability of the BBB itself is modified. This has been demonstrated with rats<br />

bearing an experimental allergic encephalomyelitis after intravenous injection. 167<br />

PEG-coated hexadecylcyanoacrylate nanospheres synthesized <strong>by</strong> Brigger et al. 168 showed a<br />

significant accumulation after intravenous injection in gliosarcoma in rats. The radioactive [ 14 C]poly(hexadecylcyanoacrylate)<br />

([ 14 C]-PHDCA) and [ 14 C]-poly(MePEG2000cyanoacrylate-cohexadecylcyanoacrylate)<br />

1:4 (PEG–PHDCA) ([ 14 C]-PEG–PHDCA) nanospheres were prepared<br />

using the nanoprecipitation technique. This procedure involves the dissolution of the radioactive<br />

polymer in a warm acetone, followed <strong>by</strong> the addition of the acetonic solution into the aqueous<br />

phase, resulting in the precipitation of nanospheres. The [ 14 C]-PHDCA and [ 14 C]-PEG–PHDCA<br />

nanospheres possessed a mean diameter of about 150 nm. Intravenous injection of the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 275<br />

[ 14 C]-PEG–PHDCA nanospheres into the rats inoculated intracerebrally with 9L gliosarcoma cells<br />

resulted in a long circulation time in the blood. Both [ 14 C]-PHDCA and [ 14 C]-PEG–PHDCA<br />

nanospheres were able to selectively extravasate the BBB and accumulated more in the gliosarcoma<br />

than in the peritumoral brain. However, this effect was significantly higher (3.1-times greater<br />

accumulation (p!0.05)) for [ 14 C]-PEG–PHDCA nanospheres than for the [ 14 C]-PHDCA nanospheres.<br />

Intravenous injection of a hydrophilic tracer [ 3 H]-sucrose to the tumor-bearing rats (seven<br />

days after the inoculation of tumor cells) to assess the permeability of the BBB showed that the 9L<br />

gliosarcoma and, to a lesser extent, the peritumoral brain region were hyperpermeable to sucrose,<br />

indicating a selective disruption of the BBB at the pathological site and slight edematous<br />

surrounding the brain. The permeability of the BBB was not observed at the intracerebral injection<br />

site of the control rats, indicating that the BBB alteration was only due to the presence of<br />

cancerous cells.<br />

15.2.6.5 Drug Delivery to Breast Cancer<br />

Resistance to cancer chemotherapy involves both altered drug activity at the designated target and<br />

modified intra-tumor pharmacokinetics (e.g., uptake and metabolism). Two proteins in particular—<br />

P-glycoprotein (P-gp) and MDR-associated protein (MRP2)—are responsible for MDR associated<br />

with a variety of cancers. 169 BCRP (breast cancer resistant protein) is another type of protein that<br />

appears to play a major role in the MDR phenotype of a specific human breast cancer. 170 Cancer<br />

cells overexpressing these proteins are resistant to several chemotherapeutic agents, including the<br />

anthracyclines, and show a reduced nuclear accumulation of anthracyclines. 171 The ability of<br />

PACA nanoparticles to overcome the MDR of cancer cells has been well documented. 137,172<br />

MRP1 is considered to be a transporter of some conjugated organic anions and drug-glutathione<br />

conjugates. 173 When delivered through PACA nanoparticles, doxorubicin forms an ion pair with<br />

the polycyanoacrylic acid and penetrates the cell membrane in resistant cells. 141 The efficacy of<br />

doxorubicin-loaded PIHCA nanoparticles was evaluated in resistant human breast adenocarcinoma<br />

cells, i.e., MCF7/VP and MCF7/doxorubicin overexpressing MRP1 and P-glycoprotein, respectively.<br />

171 The MCF7/VP and MCF7/doxorubicin were 34 and 47 times resistant to doxorubicin<br />

alone, and 2.5 and 7.7 times resistant to nanoparticle formulation, respectively, compared to MCF7<br />

cells. The nuclear accumulation of doxorubicin at 6 h after treatment was higher in MCF7 cells<br />

(133G22 and 120G41 mM when administered as free doxorubicin and nanoparticle formulation,<br />

respectively), followed <strong>by</strong> MCF7/VP (97G19 and 107G15 mM administered as free doxorubicin<br />

and nanoparticle formulation, respectively), and MCF7/doxorubicin (59G12 and 66G12 mM<br />

administered as free doxorubicin and nanoparticle formulation, respectively). At 12 h post incubation,<br />

the nuclear accumulation of doxorubicin in MCF7 cells did not increase significantly,<br />

whereas the accumulation was significantly higher in MCF7/VP cells (76G10 and 118G30 mM<br />

for free doxorubicin and nanoparticle formulation, respectively). At 18 h post incubation, the<br />

nuclear accumulation still increased in MCF7/VP cells (102G20 and 158G16 mM for free doxorubicin<br />

and nanoparticle formulation, respectively). However, the nuclear accumulation for MCF7/<br />

doxorubicin cells did not vary much with time. The confocal microscopic studies indicated the<br />

localization of doxorubicin in the nucleus of MCF7 cells, whereas the localization was observed in<br />

the cytoplasm alone in MCF7/doxorubicin cells after treatment with free doxorubicin and nanoparticle<br />

formulation. In contrast, for MCF7/VP cells the localization was in both the nucleus and the<br />

cytoplasm in cells treated with nanoparticle formulation; it was only in the cytoplasm after<br />

doxorubicin treatment.<br />

It was hypothesized that after hydrolysis of the lateral ester bonds, the ion pair passively diffuse<br />

through the lipid bilayer of the cell membrane and some amount of the polycyanoacrylic acid<br />

(3–4 carboxylic group) could interact with glutathione and create a polycyanoacrylic acid–glutathione<br />

conjugate that would inhibit the efflux of MRP1–doxorubicin. The ion pair formation<br />

between doxorubicin and polyisohexylxyanoacrylate nanoparticles could also prevent the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


276<br />

formation of the doxorubicin–glutathione–S-conjugate and hence the MRP1-mediated transport of<br />

the drug. 171 Thus, the doxorubicin-loaded PACA nanoparticles overcome the MDR in cells overexpressing<br />

P-glycoprotein and MRP1 accompanied <strong>by</strong> the nuclear accumulation of doxorubicin.<br />

In support of the above results, Kisara et al. 174 observed that the two compounds buthionine<br />

sulfoximine (BSO) and cepharanthine (CE), which decrease the glutathione content in the MDR<br />

cells, showed greater reversal of multi-drug resistance to Dox in MDR cells. Treatment of the MDR<br />

cells with BSO resulted in an enhancement of the cytotoxic effect of Dox <strong>by</strong> 1.8-fold, whereas CE<br />

caused a greater reversal of drug resistance. BSO treatment also resulted in the decrease in GSH<br />

content of MDR cells compared to that of the sensitive ones. The combination of BSO with CE<br />

caused further potentiation of the antiproliferative effect of Dox in MDR cells.<br />

15.2.6.6 Drug Delivery to Lymphatic Carcinomas<br />

The lymphatic system is the physiological system that maintains the body water balance. It also<br />

controls certain immunological responses. At times, this system acts as a medium for the mestastais<br />

of cancer cells. Due to the peculiarity of the anatomy of the lymphatic interstitium, the achievement<br />

of drug localization into the lymphatic system is limited. Much effort has been made to achieve<br />

lymphatic targeting of drugs using colloidal carriers such as biodegradable nanoparticulate systems,<br />

including nanospheres, emulsions, and liposomes. 175 The major purpose of lymphatic targeting is<br />

(1) to provide an effective anti-cancer chemotherapy that will prevent the metastasis of tumor cells<br />

<strong>by</strong> accumulating the drug in the regional lymph node, and (2) to enhance the localization of<br />

diagnostic agents in the regional lymph node to visualize the lymphatic vessels before surgery.<br />

In the early days of lymphatic-targeting research, some success was achieved in delivering the<br />

emulsions and liposomes to the lymphatic system after intramuscular 176 and subcutaneous routes of<br />

administration. 177 For effective penetration into the lymphatic interstitium, the carrier systems need<br />

to possess a smaller size, hydrophobicity, and a negative surface charge. 178 PIBCA nanocapsules<br />

showed enhanced lymphatic targeting and regional lymph node retention of 12-(9-anthroxy) stearic<br />

acid, a lipophilic model compound. 175 The route of administration of the delivery system has a<br />

greater contribution in the delivery of nanoparticles to the lymphatic tumor in mice. 179 In vivo<br />

studies conducted <strong>by</strong> Reddy et al. 74 on mice bearing Dalton’s Lymphoma tumor have showed that<br />

99m Tc-Dox-loaded PBCA nanoparticles exhibited enhanced accumulation in the lymph node,<br />

compared with the 99m Tc-Dox, after subcutaneous injection. A significantly high tumor uptake<br />

of DPBC nanoparticles (w13-fold higher at 48 h post injection) (P!0.001) was found, compared<br />

to free doxorubicin. The accumulation of the nanoparticles in the tumor increased with time,<br />

indicating their slow penetration from the injection site into the tumor. Such a nanoparticulate<br />

system would be advantageous in the lymphoma treatment, offering drug targeting and the<br />

prolonged availability of the drug to the tumor.<br />

15.2.6.7 Drug Delivery to Gastric Carcinomas<br />

Nanotechnology for Cancer Therapy<br />

Gastric carcinoma falls into the category of the difficult-to-treat tumors. Though the PACA nanoparticles<br />

are considered most promising in the tumor-treatment applications, their use in the<br />

delivery of anti-cancer agents to gastric tumors remains limited.<br />

Magnetically targeted systems can be classified in the new generation of drug carriers.<br />

The advantage of these carriers is their ability to minimize the uptake <strong>by</strong> the reticuloendothelial<br />

system. 180 Intra-arterial administration of these systems in the form of magnetic albumin microspheres<br />

181 and magnetic liposomes 182 has been shown to treat the tumors successfully under the<br />

influence of a strong magnetic field. Recently, Gao et al. 183 prepared aclacinomycin-loaded<br />

magnetic iron oxide–PBCA nanoparticles using the interfacial polymerization technique. The<br />

technique involved dispersion of the drug in diluted hydrochloric acid into magnetic fluid, and<br />

the successive addition of the mixture to a hexane containing span 80 and polysorbate 80.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 277<br />

The buytlcyanoacrylate monomer was added into the above dispersion under stirring, and after 6 h<br />

the newly formed magnetic nanoparticles were separated using a magnet, washed with methanol<br />

and water, lyophilized, and sterilized <strong>by</strong> 60 Co irradiation (15 kGy). These magnetic particles were<br />

210 nm in diameter, and they revealed a core-shell structure in which the iron oxide core was<br />

covered <strong>by</strong> the polymer coat. The aclacinomycin-loaded nanoparticles were evaluated in mice<br />

implanted with gastric carcinoma near the right fore feet. Administration of drug-loaded magnetic<br />

nanoparticles <strong>by</strong> intravenous injection into the tumor-bearing mice implanted with a magnet in the<br />

tumor resulted in a higher tumor-inhibition rate than the free aclacinomycin. The nanoparticle<br />

formulation greatly reduced the tumor mass compared to the free aclacinomycin and the drugfree<br />

carrier. The drug-loaded magnetic nanoparticles showed similar activity to that of the drug-free<br />

magnetic nanoparticles in vitro in the absence of a magnetic field in the gastric cancer cell line<br />

MKN-45. They did, however, show enhanced antitumor activity in vivo in tumor-bearing mice.<br />

15.2.6.8 Chemoembolization Using Alkylcyanoacrylates<br />

HCC is the most common liver tumor, with heterogeneity in the tumor behavior and the underlying<br />

liver disease. Recent combinations such as cisplatin, interferon, Adriamycin, and 5-FU are extremely<br />

toxic and yield response rates of only 20%, with no survival advantage compared to supportive care<br />

alone. 184 Higher concentrations of cancer chemotherapeutic agents can be delivered directly to the<br />

HCC via the hepatic arterial route. Considering that this route is the major vascular supply of these<br />

tumors, an even larger number of papers have reported the experience of hepatic artery<br />

chemotherapy or hepatic artery chemoembolization (TACE) with single agents, or with a dizzying<br />

combination of agents, and at doses not replicated <strong>by</strong> any two institutions. Loewe et al. 185 evaluated<br />

the potential of transarterial permanent embolization with the use of a mixture of cyanoacrylate and<br />

lipiodol for the treatment of unresectable primary HCC. Loewe et al. 186 used NBCA for hepatic<br />

artery embolization for the treatment of small-bowel neuroendocrine metastases to the liver. The<br />

results revealed that the permanent embolization of hepatic arteries as part of a multimodality<br />

treatment protocol is beneficial in long-term follow-up for patients with metastasized small-bowel<br />

neuroendocrine tumors. The use of cyanoacrylate is safe and effective as an embolic agent.<br />

Transarterial embolization (TAE) with the use of microspheres and Lipiodol and cyanoacrylate<br />

for unresectable HCC is a feasible treatment modality. A retrospective analysis of 46 patients with<br />

histologically confirmed HCC was made who were treated with TAE of the hepatic arteries. 187 To<br />

induce permanent embolization, the microspheres (Embosphere; 100–700 mm) and a mixture of<br />

ethiodized oil (Lipiodol Ultrafluide) with cyanoacrylate (Glubran) was administrated. No patient<br />

died during embolization or within the first 24 h. Severe procedure-related complications were<br />

observed in 2 patients. At the time of the analysis, 38 of 46 patients were alive. The 180-, 360-, 520-<br />

, and 700-day cumulative survival rates for the total study population were 80.6%, 70.7%, 70.7%,<br />

and 47.1%, respectively, with a median survival of 666 days.<br />

A procedure for effective and promising preoperative embolization of carotid body tumors was<br />

reported <strong>by</strong> Harman et al. 188 Ultrasound-guided direct percutaneous injection of n-butyl cyanoacrylate<br />

was given, and angiographic road map assistance was used for protection of parent arteries<br />

during the injection. After embolization, complete devascularization of the tumor was achieved<br />

without complications. The tumor was removed surgically with minimal blood loss. Transcatheter<br />

arterial embolization (TAE) of splanchnic arterial branches to allow continuous application of<br />

repeat hepatic arterial infusion chemotherapy (HAIC) was assessed. 189 One hundred and twentyeight<br />

patients with unresectable advanced liver cancer were implanted with a percutaneous port<br />

catheter system and TAE of splanchnic arteries with coils and/or NBCA. The recanalization rate<br />

between coil-embolized and NBCA- or NBCA-coil-embolized arteries, and frequency of heterogeneously<br />

poor distribution was compared between patients with single arteries and those with<br />

multiple hepatic arteries. The arteries once embolized with coils alone spontaneously recanalized at<br />

a significantly higher rate than those with NBCA. A hepatic artery embolization study carried out<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


278<br />

<strong>by</strong> Loewe et al. 186 using NBCA and ethiodized oil for the treatment of small-bowel neuroendocrine<br />

metastases to the liver for the treatment of liver metastases from neuroendocrine<br />

small-bowel tumors also concluded that the use of cyanoacrylate as an embolic agent is safe<br />

and effective.<br />

The potential of transarterial permanent embolization with the use of a mixture of cyanoacrylate<br />

and lipiodol for treatment of unresectable primary HCC was also assessed <strong>by</strong> Loewe et<br />

al. 185 The study included 36 patients with histologically proven HCC who were treated with<br />

transarterial embolization of the hepatic arteries. The study indicated that TAE with use of<br />

cyanoacrylate and lipiodol for unresectable HCC is a feasible treatment modality. A similar<br />

study conducted <strong>by</strong> Berghammer et al. 190 confirmed the safety of the procedure with minimum<br />

side effects. Thus, it constitutes a valuable therapeutic option for patients with Okuda stage I<br />

and II HCC.<br />

A right gastric artery (RGA) embolization study to prevent acute gastric mucosal lesions caused<br />

<strong>by</strong> an influx of anti-cancer agents into the RGA in patients undergoing repeat HAIC was conducted<br />

on 217 patients with malignant hepatic tumors 191 using metallic coils and/or a mixture of n-butyl<br />

cyanoacrylate (n-BCA). RGA embolization was technically successful in the majority of patients<br />

(93%), with the lowest incidence of major complications. The clinical experience of Nadalini<br />

et al. 192 using isobutyl-2-cyanoacrylate vesical for embolization of the hypogastric arteries in<br />

cases of serious hemorrhage of the bladder and prostate is also reported. The effect was immediate<br />

results in the majority of cases, a decidedly positive outcome, especially considering the serious<br />

conditions of certain neoplastic patients. Isobutyl-2-cyanoacrylate suspension in Lipiodol was also<br />

used to treat percutaneous transcatheter embolization of the renal artery in clear cell carcinoma 193<br />

in nine patients. In most of the patients, the procedure was found to be palliative, with no complications<br />

attributed to the glue or to oil emboli. Preoperative embolization with isobutyl-2cyanoacrylate<br />

<strong>by</strong> means of an intra-arterial catheter during selective angiography was adopted<br />

<strong>by</strong> Carmignani et al. 194 in two cases of carcinoma of the kidney.<br />

Traditional preoperative embolization via a transarterial approach has proved beneficial, but it<br />

is often limited <strong>by</strong> complex vascular anatomy and unfavorable locations. 195 Paragangliomas, or<br />

glomus tumors, are the neoplasms of the head and neck. They are remarkably vascular, so surgical<br />

resection can be complicated <strong>by</strong> rapid and dramatic blood loss. 196 These tumors can develop in the<br />

middle ear (glomus tympanicum), the jugular foramen of the skull base (glomus jugulare), or the<br />

head and neck area (glomus caroticum, glomus vagale). Surgical removal of these tumors is also<br />

often associated with a significant intraoperative bleeding rate because of their vascular nature. 197–199<br />

Direct percutaneous injection of n-butyl cyanoacrylate resulted in the effective devascularization<br />

of craniofacial tumors 200 and the embolization of oral tumors. 201 However, the technique<br />

involved additional risks and was not widely adopted. In a study in human patients <strong>by</strong> Abud<br />

et al., 196 the presurgical devascularization was achieved <strong>by</strong> placing the diagnostic catheter in<br />

the common carotid artery to guide the puncture and perform the control angiography during<br />

and after the injection of the cyanoacrylate. The percutaneous devascularization of head<br />

and neck paragangliomas through the intralesional injection of cyanoacrylate resulted in effective<br />

devascularization. It could be a safe and effective technique for managing such<br />

clinical lesions.<br />

15.2.6.9 Delivery of Peptide Anti-Cancer Drugs<br />

Nanotechnology for Cancer Therapy<br />

Ras mutations exist in almost 20–30% of human tumors, and the ras oncogenes possess single-point<br />

mutations in the sequence coding for the active site of RAS protein at codons 12, 13b, and 61. 202<br />

These point mutations are good targets for the antisense oligonucleotides and can suppress the<br />

translation and targeted mutant mRNA. The mutated ras genes are involved in the cell proliferation<br />

and tumorigenicity. Antisense oligonucleotides (ODN) are molecules that are able to inhibit oncogene<br />

expression, being therefore potentially active for the treatment of viral infections or cancer.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 279<br />

However, because of their poor stability in biological media and their weak intracellular<br />

penetration, colloidal drug carriers such as nanoparticles were employed for the delivery of<br />

oligonucleotides. 203 Association with nanoparticles protects the ODN against degradation and<br />

enables them to penetrate more easily into different types of cells. Thus, they were shown to<br />

improve the efficiency in inhibiting the proliferation of cells expressing the point-mutated Ha-ras<br />

gene. In vivo, the PACA nanoparticles were able to efficiently distribute the ODNs to the liver. As<br />

the ODNs have no affinity toward the PACA structure, the ion pair technique was adopted, using<br />

cationic surfactants such as cetyltrimethyl ammonium bromide or diethylaminoethyl dextran<br />

adsorbed onto the particle surface. 204,205 Spontaneous interpolymer complexation between<br />

cationic polyelectrolytes and DNA is known and is largely the result of cooperative electrostatic<br />

forces. 206<br />

ODNs adsorbed onto the PIHCA nanoparticles inhibited the Ha-Ras dependant tumor growth in<br />

mice. The ODN-nanoparticle formulation was tested in two cell lines, HBL100rasl and<br />

HBL100neo. 202 HBL100rasl is a clone obtained from the human mammary cell line. It expresses<br />

normal Ha-ras and Ha-ras carrying the G/U point mutation in codon 12 coding for Val instead of<br />

Gly at position 12. HBL100neo is a clone transformed only with the pSV2 vector, and it expresses<br />

only normal Ha-ras. Three 12-mer oligonucleotides—(1) an antisense oligonucleotide (AS-Val, 5 0 -<br />

CACCGACGGCGC-3 0 ) directed against and centered at the point mutation in codon 12 of the Ha-ras<br />

mRNA, (2) an antisense oligonucleotide (AS-Gly, 5 0 -CACCGCCGGCGC-3 0 )targetedtothe<br />

equivalent sequence of the normal Ha-ras mRNA, and (3) the 5 0 /3 0 inverted sequence of AS-Val<br />

that contains the same bases as the antisense sequence but in reverse orientation (INV-Val, 5 0 -<br />

CGCCGGAGCCAC-3 0 ). The inhibitory effect of AS-Val adsorbed on the nanoparticles was<br />

found to be at a concentration 100-times lower than that for the free AS-Val when tested on<br />

HBL100rasl. Cellular uptake studies using 32 P-labeling of free ODN and ODN adsorbed nanoparticle<br />

formulation revealed that the ODNs adsorbed onto the nanoparticles, incorporated at a lower<br />

rate than the free ODN but remained intact in the cells even after 72 h post incubation; the free ODN,<br />

in contrast, degraded after 3 h. Upon administration into the nude mice treated with HBL100rasl<br />

cells, the AS-Val nanoparticle formulation showed more potential tumor growth inhibition<br />

compared to the INV-Val nanoparticle formulation, indicating the sequence-dependant effect of<br />

ODN on the tumor-inhibition properties.<br />

The cationic surfactants used for the preparation of ODN–PACA nanoparticles may cause<br />

toxicity. In addition, the esterase-induced erosion of PACAs causes rapid desorption of ODNs in<br />

serum. 207 To overcome these drawbacks, the PACA nanocapsules with aqueous core have been<br />

designed. 208,209 These nanocapsules enhanced the protection and serum stability of the ODNs. 209<br />

Intratumoral injection of PACA nanoparticles containing Phosphorothioate oligonucleotides<br />

directed against EWS Fli-1 chimeric RNA led to a significant tumor growth inhibition compared<br />

to the free ODN in mice-bearing Ewing sarcoma.<br />

Peptides like Leu-enkephalin dalargin and the Met-enkephalin kyotorphin normally do not<br />

cross the BBB when given systemically. Transport of these neuropeptides across the BBB was<br />

achieved <strong>by</strong> adsorption onto the surface of PBCA nanoparticles and successive coating of the<br />

nanoparticles with polysorbate (Tween) 80. 210<br />

Birrenbach and Speiser 5 described the polymerization process for the preparation of hydrophilic<br />

micelles containing solubilized drug molecules including labile proteins in a colloidal<br />

aqueous system of dissolved monomers. The hydrocarbon medium constituted the outer phase.<br />

After secondary solubilization with the aid of selected surfactants, the polymerization of micelles<br />

under different conditions was performed. The entrapped tagged material (human 125I-immunoglobulin<br />

G) showed a stable fixation in the nanoparticles during long-term in vitro liberation trials.<br />

Nanoparticles were spherical in shape, 80 nm in diameter, and embedded with antigenic material<br />

(tetanus toxoid and human immunoglobulin G) for parenteral use. These preparations showed intact<br />

biological activity and high antibody production in animals.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


280<br />

15.3 CONCLUSIONS<br />

Drug delivery to tumors has recently been the area of significant interest in the field of drug<br />

delivery. Nanoparticles, due to their many inherent advantages, gained an important place in this<br />

area of research. Though the alkylcyanoacrylates were initially used as tissue adhesives, the<br />

discovery of the simple anionic polymerization of alkylcyanoacrylates has renewed interest in<br />

their applicability to drug-delivery research. Poly(alkyl cyanoacrylates) with a higher alkyl chain<br />

length are relatively less toxic than those with a lower alkyl chain length. Several interesting studies<br />

were reported on the usefulness of poly(alkyl cyanoacrylates) for delivering anti-cancer drugs to<br />

different types of cancers, including breast cancer and gastric cancer. Interesting results were<br />

observed in the case of delivery of doxorubicin-loaded poly(alkyl cyanoacrylate) nanoparticles<br />

for the treatment of liver tumors. Though the initial results on the delivery of doxorubicin-loaded<br />

poly-butylcyanoacrylate nanoparticles to lymphoma are encouraging, more experimental evidence<br />

is needed to evaluate their efficacy on tumors. The PEG modification of alkylcyanoacrylates has<br />

improved the delivery of their nanoparticles to the brain. Cellular resistance to multiple drugs<br />

represents a major problem in cancer chemotherapy. This drug resistance may appear clinically<br />

either as a lack of tumor size reduction or as the occurrence of clinical relapse after an initial<br />

positive response to treatment with antitumor agents. The resistance mechanism can have different<br />

origins, either directly linked to specific mechanisms developed <strong>by</strong> the tumor tissue or connected to<br />

the more general problem of distribution of a drug toward its targeted tissue. The role of P-glycoprotein<br />

in decreasing the uptake of anti-cancer drugs in multi-drug resistant (MDR) tumors through<br />

efflux is well studied. Reports on the importance of nanoparticles in reversion of multi-drug<br />

resistance mediated <strong>by</strong> P-glycoprotein in MDR tumors reveal the potential of using PBCA nanoparticles<br />

in MDR tumors. A commercial application of such a strategy is on the way: doxorubicinloaded<br />

poly(alkyl cyanoacrylate) nanoparticles have reached phase-II clinical trials for the treatment<br />

of resistant tumors. Several reports have elucidated the mechanism of nanoparticle uptake <strong>by</strong><br />

the cells, but the intracellular trafficking and fate of these nanoparticles in the cell remain unclear.<br />

Of late, poly(alkyl cyanoacrylate) nanoparticles have also been considered as a potential system for<br />

delivering proteins and peptide drugs to tumors.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Fenton, R. G. and Longo, D. I., Cell biology of cancer, In Harrison’s Internal Medicine, Fauci, A. S.,<br />

Braunwald, E., Isselbacher, K. J., Wilson, J. D., Martin, J. B., Kasper, D. L., Hauser, S. L., and<br />

Longo, D. I., Eds., vol. 1, 14 th ed., McGraw Hill, New York, p. 505, 1998.<br />

2. Reddy, L. H., Drug delivery to tumors: Recent strategies, J. Pharm. Pharmacol., 57, 1231, 2005.<br />

3. Williams, J. et al., Nanoparticle drug delivery system for intravenous delivery of topoisomerase<br />

inhibitors, J. Control. Release, 91, 167, 2003.<br />

4. De Jaeghere, F., Doelker, E., and Gurny, R., Nanoparticles, In Encyclopedia of Controlled Drug<br />

Delivery, Mathiowitz, E., Ed., Wiley, New York, p. 641, 1999.<br />

5. Birrenbach, B. and Speiser, P. P., Polymerized micelles and their use as adjuvants in immunology,<br />

J. Pharm. Sci., 65, 1763, 1976.<br />

6. Rhodes, C. T., Wai, K., and Banker, G. S., Molecular scale drug entrapment as a precise method of<br />

controlled drug release. II. Facilitated drug entrapment to polymeric colloidal dispersions, J. Pharm.<br />

Sci., 59, 1578, 1970.<br />

7. Boylan, J. C. and Banker, G. S., Molecular scale drug entrapment as a precise method of controlled<br />

drug release. IV. Entrapment of anionic drugs <strong>by</strong> polymer gelation, J. Pharm. Sci., 62, 1177, 1973.<br />

8. Langer, R., Biomaterials in drug delivery and tissue engineering: one laboratory’s experience, Acc.<br />

Chem. Res., 33, 94, 2000.<br />

9. Lanza, R. P., Langer, R., and Chick, W. L., Principles of Tissue Engineering, Academic Press,<br />

Austin, TX, p. 405, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 281<br />

10. Gilbert, R. G., Emulsion Polymerization: A Mechanistic Approach, Academic Press, London, p. 30,<br />

1995.<br />

11. Bovey, F. A. et al., Emulsion Polymerization, vol. 1, Interscience Publishers, New York, p. 15, 1955.<br />

12. Ballard, M. J., Napper, D. H., and Gilbert, R. G., Kinetics of emulsion polymerization of methyl<br />

methacrylate. Polymer chemistry special issue, J. Polym. Sci., 22, 3225, 1984.<br />

13. Ljungstedt, I., Ekman, B., and Sjoholm, I., Detection and separation of lymphocytes with specific<br />

surface receptors <strong>by</strong> using microparticles, Biochem. J., 170, 161, 1978.<br />

14. Krause, H. J., Schwarz, A., and Rohdewald, P., Interfacial polymerization, a useful method for the<br />

preparation of polymethylcyanoacrylate nanoparticles, Drug Dev. Ind. Pharm., 12, 527, 1986.<br />

15. Radwan, M. A., Preparation and in vivo evaluation of parenteral metoclopramide-loaded poly(alkylcyanoacrylate)<br />

nanospheres in rats, J. Microencapsul., 18, 467, 2001.<br />

16. Couvreur, P. et al., Polycyanoacrylate nanocapsules as potential lysosomotropic carriers: Preparation,<br />

morphological and sorptive properties, J. Pharm. Pharmacol., 31, 331, 1979.<br />

17. Grislain, L. et al., Pharmacokinetics and biodistribution of a biodegradable drug-carrier, Int.<br />

J. Pharm., 15, 335, 1983.<br />

18. Behan, N., Birkinshaw, C., and Clarke, N., Poly n-butyl cyanoacrylate nanoparticles: a mechanistic<br />

study of polymerization and particle formation, Biomaterials, 22, 1335, 2001.<br />

19. Seijo, B. et al., Design of nanoparticles of less than 50 nm diameter: Preparation, characterization<br />

and drug loading, Int. J. Pharm., 62, 1, 1990.<br />

20. Reddy, L. H. and Murthy, R. S. R., Influence of polymerization technique and experimental variables<br />

on the particle properties and release kinetics of methotrexate from poly(butylcyanoacrylate) nanoparticles,<br />

Acta Pharm., 54, 103, 2004.<br />

21. Reddy, L. H. and Murthy, R. S. R., Study of influence of polymerization factors on formation of<br />

poly(butyl cyanoacrylate) nanoparticles and in vitro drug release kinetics, ARS Pharm., 45, 211,<br />

2004.<br />

22. Fallaouh, N. A. K. et al., Development of a new process for the manufacture of polyisobutylcyanoacrylate<br />

nanocapsules, Int. J. Pharm., 28, 125, 1986.<br />

23. Florence, A. T., Whateley, T. L., and Wood, D. A., Potentially biodegradable microcapsules with<br />

poly-2-cyanoacrylate membranes, J. Pharm. Pharmacol., 31, 422, 1979.<br />

24. Chouinard, F. et al., Preparation and purification of polyisohexylcyanoacrylate nanocapsules, Int.<br />

J. Pharm., 72, 211, 1991.<br />

25. Palumbo, M. et al., Improved antioxidant effect of idebenone-loaded polyethyl-2-cyanoacrylate<br />

nanocapsules tested on human fibroblasts, Pharm. Res., 19, 71, 2002.<br />

26. Kopf, H. et al., Studium der Mizellpolymerisation in Gegenwart niedermolekularer Arzneistoffe. I.<br />

Herstellung und Isolierung der Nanopartikel, Restmonomerenbestimmung, physikalischchemische<br />

Daten, Pharm. Ind., 38, 281, 1976.<br />

27. Laakso, T. and Stjarnkvistjoholm, I., Biodegradable microspheres. VI. Lysosomal release of covalently<br />

bound antiparasitic drugs from starch microparticles, J. Pharm. Sci., 76, 134, 1987.<br />

28. Beck, P. et al., Influence of polybutyl cyanoacrylate nanoparticles and liposomes on the efficacy and<br />

toxicity of the anticancer drug mitoxantrone in murine tumor models, J. Microencapsul., 10, 101,<br />

1993.<br />

29. Illum, L. et al., Evaluation of carrier capacity and release characteristic for poly(butyl-2-cyanoacrylate)<br />

nanoparticles, Int. J. Pharm., 30, 17, 1986.<br />

30. Couvreur, P. et al., Adsorption of antineoplastic drugs to polyalkyl cyanoacrylate nanoparticles and<br />

their release in calf serum, J. Pharm. Sci., 68, 1521, 1979.<br />

31. Kreuter, J. and Hartman, H. R., Comparative study on the cytostatic effects and tissue distribution of<br />

5-fluorouracil in a free from and bound to polybutylcyanocrylate nanoparticles in sarcoma 180bearing<br />

mice, Oncology, 40, 363, 1983.<br />

32. Marchal-Heussler, I. et al., Antiglaucomatous activity of betaxolol chlorhydrate sorbed onto<br />

different isobutyl cyanoacrylate nanoparticle preparations, Int. J. Pharm., 58, 115, 1990.<br />

33. Brasseur, N., Brault, D., and Couvreur, P., Adsorption of hematoporphyrin onto polyalkyl cyanoacrylate<br />

nanoparticles: carrier capacity and drug release, Int. J. Pharm., 70, 129, 1991.<br />

34. Gaspar, R., Preat, V., and Roland, M., Nanoparticles of polyisohexyl cyanoacrylate (PIHCA) as<br />

carriers of primaquine: formulation, physico-chemical characterization and acute toxicity, Int.<br />

J. Pharm., 68, 111, 1991.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


282<br />

Nanotechnology for Cancer Therapy<br />

35. Verdun, C. et al., Development of a nanoparticle controlled-release formulation for human use,<br />

J. Control. Release, 3, 205, 1986.<br />

36. Harmia, T., Speiser, P., and Kreuter, J., Optimization of pilocarpoine loading onto nanoparticles <strong>by</strong><br />

sorption procedures, Int. J. Pharm., 33, 45, 1986.<br />

37. Poupaert, J. H. and Couvreur, P. A., Computationally derived structural model of doxorubicin<br />

interacting with oligomeric polyalkylcyanoacrylate in nanoparticles, J. Control. Release, 92, 19,<br />

2003.<br />

38. Chauvierre, C. et al., Heparin coated poly(alkylcyanoacrylate) nanoparticles coupled to hemoglobin:<br />

A new oxygen carrier, Biomaterials, 25, 3081, 2004.<br />

39. Simeonova, M. et al., Study on the role of 5-fluorouracil in the polymerization of butylcyanoacrylate<br />

during the formation of nanoparticles, J. Drug Target., 12, 49, 2004.<br />

40. Stella, B. et al., Design of folic acid-conjugated nanoparticles for drug targeting, J. Pharm. Sci., 89,<br />

1452, 2000.<br />

41. Kreuter, J., Physicochemical characterization of polyacrylic nanoparticles, Int. J. Pharm., 14,<br />

43, 1983.<br />

42. Kreuter, J. et al., Polybutyl cyanoacrylate nanoparticles for the delivery of [ 75 SE]norcholesterol, Int.<br />

J. Pharm., 16, 105, 1983.<br />

43. Magenheim, B. and Benita, S., Nanoparticle characterization: A comprehensive physicochemical<br />

approach, STP Pharma Sci., 1, 121, 1991.<br />

44. Edstrom, R. D. et al., Viewing molecules with scanning tunneling and atomic force microscopy,<br />

FASEB J., 4, 3144, 1990.<br />

45. Zasadzinski, J. A. N., Scanning tunneling microscopy with applications to biological surfaces,<br />

Biotechniques, 7, 174, 1989.<br />

46. Fuchs, H. and Laschinski, R., Surface investigations with a combined scanning-electron scanning<br />

tunneling microscope, Scanning, 12, 126, 1990.<br />

47. Gedde, U. W., Thermal analysis of polymers, Drug Dev. Ind. Pharm., 16, 2465, 1990.<br />

48. Muller, R. H. et al., Alkyl cyanoacrylate drug carriers. I. Physicochemical characterization of<br />

nanoparticles with different chain length, Int. J. Pharm., 84, 1, 1992.<br />

49. Andreade, J. D. et al., Contact angles at the solid–water interface, J. Colloid Interface Sci., 72, 488,<br />

1979.<br />

50. Wilkins, D. J. and Myers, P. A., Studies on the relationship between the electrophoretic properties of<br />

colloids and their blood clearance and organ distribution in the rat, Br. J. Exp. Pathol., 47, 568, 1966.<br />

51. Baszkin, A. and Lyman, D. J., The interaction of plasma proteins with polymers. I. Relationship<br />

between polymer surface energy and protein adsorption/desorption, J. Biomed. Mater. Res., 14, 393,<br />

1980.<br />

52. Lindsay, R. M. et al., Blood surface interactions, Trans. Am. Soc. Artif. Intern. Organs, 26, 603,<br />

1980.<br />

53. Andreade, J. D. et al., Surface characterization of poly(hydroxyethyl methacrylate) and related<br />

polymers. I. Contact angle methods in water, J. Polym. Sci. Polym. Symp., 66, 131, 1979.<br />

54. El-Egakey, M. A., Bentele, V., and Kreuter, J., Molecular weights of polycyanoacrylate nanoparticles,<br />

Int. J. Pharm., 13, 349, 1983.<br />

55. Muller, R. H. et al., In vitro model for the degradation of alkylcyanoacrylate nanoparticles, Biomaterials,<br />

11, 590, 1990.<br />

56. Scherer, D., Robinson, J. R., and Kreuter, J., Influence of enzymes on the stability of polybutylcyanoacrylate<br />

nanoparticles, Int. J. Pharm., 101, 165, 1994.<br />

57. Sommerfeld, P., Schroeder, U., and Sabel, B. A., Long-term stability of PBCA nanoparticle suspensions<br />

suggests clinical usefulness, Int. J. Pharm., 155, 201, 1997.<br />

58. Maincent, P. et al., Disposition kinetics and oral bioavailability of vincamine-loaded polyalkyl<br />

cyanoacrylate nanoparticles, J. Pharm. Sci., 75, 955, 1986.<br />

59. Kreuter, J., Peroral administration of nanoparticles, Adv. Drug Deliv. Rev., 7, 71, 1991.<br />

60. Damge, C. et al., New approach for oral administration of insulin with polyalkylcayanoacrylate<br />

nanocapsules as drug carrier, Diabetes, 37, 246, 1988.<br />

61. Damge, C. et al., Nanocapsules as carriers for oral peptide delivery, J. Control. Release, 13, 233,<br />

1990.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 283<br />

62. Wade, C. and Leonard, F., Degradation of poly-(methyl-2-cyanoacrylates), J. Biomed. Mater. Res.,<br />

6, 215, 1972.<br />

63. Lenaerts, V. et al., Degradation of poly(isobutylcyanoacrylate) nanoparticles, Biomaterials, 5, 65,<br />

1984.<br />

64. Douglas, S. J., Illum, L., and Davis, S. S., Particle size and size distribution of Poly(butyl-2-cyanoacrylate)<br />

nanoparticles. II. Influence of stabilizers, J. Colloid Interface Sci., 103, 154, 1985.<br />

65. Couvreur, P., Design of biodegradable polyalkylcyanoacrylate nanoparticles as a drug carrier, In<br />

Microspheres and Drug Therapy, Davis, S. S., Illum, L., McVie, J. G., and Tomlinson, E., Eds.,<br />

Elsevier, Amsterdam, p. 103, 1984.<br />

66. Washington, C., Drug release from monodisperse systems: A critical review, Int. J. Pharm., 58, 1,<br />

1990.<br />

67. Leavy, M. V. and Benita, S., Drug release from submicron O/W emulsion: a new in vitro kinetic<br />

evaluation model, Int. J. Pharm., 66, 29, 1990.<br />

68. Calvo, P., Vila-Jato, J. L., and Alonso, M. J., Comparative in vitro evaluation of several colloidal<br />

systems, nanoparticles, nanocapsules and nanoemulsions as ocular drug carriers, J. Pharm. Sci., 85,<br />

530, 1996.<br />

69. El-Samaligly, M. S., Rohdewald, P., and Mohmoud, H. A., Polyalkyl cyanoacrylate nanocapsules,<br />

J. Pharm. Pharmacol., 38, 216, 1986.<br />

70. LeRay, A. M. et al., End chain radiolabeling and in vitro stability studies of radiolabeled poly<br />

(hydroxyl) nanoparticles, J. Pharm. Sci., 83, 845, 1994.<br />

71. Barratt, G., Colloid drug carriers: achievements and perspectives, Cell. Mol. Life Sci., 60, 21, 2003.<br />

72. Simeonova, M. et al., Tissue distribution of polybutylcyanoacrylate nanoparticles loaded with spinlabelled<br />

nitrosourea in Lewis lung carcinoma-bearing mice, Acta Physiol. Pharmacol. Bulg., 20, 77,<br />

1994.<br />

73. Reddy, L. H. and Murthy, R. S. R., Pharmacokinetics and bio-distribution of doxorubicin loaded<br />

poly(butyl) cyanoacrylate nanoparticles synthesized <strong>by</strong> two different techniques, Biomed. Pap., 148,<br />

1611, 2004.<br />

74. Reddy, L. H., Sharma, R. K., and Murthy, R. S.R, Enhanced tumor uptake of doxorubicin loaded<br />

poly(butyl cyanoacrylate) nanoparticles in mice bearing dalton’s lymphoma tumor, J. Drug Target.,<br />

12, 443, 2004.<br />

75. Gipps, E. M. et al., Distribution of polyhexyl cyanoacrylate nanoparticles in nude mice bearing<br />

human osteosarcoma, J. Pharm. Sci., 75, 256, 1986.<br />

76. Illum, L. et al., Tissue distribution of poly(hexyl 2-cyanoacrylate) nanoparticles coated with monoclonal<br />

antibodies in mice bearing human tumor xenografts, J. Pharmacol. Exp. Ther., 230, 733,<br />

1984.<br />

77. Kante, B. et al., Tissue distribution of [ 3 H] actinomycin D adsorbed on polybutylcyanoacrylate<br />

nanoparticles, Int. J. Pharm., 7, 45, 1980.<br />

78. Douglas, S. J., Davis, S. S., and Illum, L., Biodistribution of poly(butyl-2-cyanoacrylate) nanoparticles<br />

in rabbits, Int. J. Pharm., 34, 145, 1986.<br />

79. Lobenberg, R. et al., Body distribution of azidothymidine bound to hexylcyanoacrylate nanoparticles<br />

after i.v. injection to rats, J. Control. Release, 50, 21, 1998.<br />

80. Borchard, G. et al., Uptake of surfactant-coated poly(methylmethacrylate)-nanoparticles <strong>by</strong> bovine<br />

brain microvessel endothelial cell monolayers, Int. J. Pharm., 110, 29, 1994.<br />

81. Gulyaev, A. E. et al., Significant transport of doxorubicin into the brain with polysorbate 80-coated<br />

nanoparticles, Pharm. Res., 16, 1564, 1999.<br />

82. Woodcock, D. M. et al., Reversal of multidrug resistance <strong>by</strong> surfactants, Br. J. Cancer, 66, 62, 1992.<br />

83. Nerurkar, M. M., Burto, P. S., and Borchardt, R. T., The use of surfactants to enhance the permeability<br />

of peptides through Caco-2 cells <strong>by</strong> inhibition of an apically polarized efflux system,<br />

Pharm. Res., 13, 528, 1996.<br />

84. Ghanem, G. E. et al., Labelled polycyanoacrylate nanoparticles for human in vivo use, Appl. Radiat.<br />

Isot., 44, 1219, 1993.<br />

85. Skidan, I. N. et al., Modification of pharmacokinetics and antitumor efficacy of doxorubicin bound to<br />

polysorbate 80-coated polybutylcyanoacrylate nanoparticles. In Third Central European Conference<br />

on Human Tumor Markers, Biomarkers Environ. J., 4, 39, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


284<br />

Nanotechnology for Cancer Therapy<br />

86. Pappo, J. and Ermak, T. H., Uptake and translocation of fluorescent latex particles <strong>by</strong> rabbit Peyer’s<br />

patch follicle epithelium: a quantitative model for M cell uptake, Clin. Exp. Immunol., 76, 144, 1989.<br />

87. Eldridge, J. H. et al., Controlled release in the gut-associated lymphoid tissues. 1. Orally administered<br />

biodegradable microspheres target the Peyer’s patches, J. Control. Release, 11, 205, 1990.<br />

88. Dembri, A. et al., Targeting of 3 0 -Azido3 0 -deoxythymidine (AZT)-loaded poly(isohexylcyanoacrylate)<br />

nanospheres to the gastrointestinal mucosa and associated lymphoid tissues, Pharm. Res., 18,<br />

467, 2001.<br />

89. Reszka, R. et al., Body distribution of free, liposomal and nanoparticle-associated mitoxantrone in<br />

B16-melanoma-bearing mice, J. Pharmacol. Exp. Ther., 280, 232, 1997.<br />

90. Fattal, E. et al., Treatment of experimental salmonellosis in mice with ampicillin-bound nanoparticles,<br />

Antimicrob. Agents Chemother., 33, 1540, 1989.<br />

91. Chiannilkulchai, N. et al., Doxorubicin-loaded nanoparticles: increased efficacy in murine hepatic<br />

metastases, Selective Cancer Ther., 5, 1, 1989.<br />

92. Storm, G. et al., Surface modification of nanoparticles to oppose uptake <strong>by</strong> the mononuclear phagocyte<br />

system, Adv. Drug Deliv. Rev., 17, 31, 1995.<br />

93. Gref, R. et al., In Microparticulate Systems for the Delivery of Proteins and Vaccines, Chen, S. and<br />

Bernstein, H., Eds., Marcel Dekker, New York, p. 279, 1996.<br />

94. Gabizon, A., Selective tumor localization and improvement of therapeutic index of anthracyclines<br />

encapsulated in long circulating liposomes, Cancer Res., 52, 891, 1992.<br />

95. Gref, R., Biodegradable long-circulating polymeric nanospheres, Science, 263, 1600, 1994.<br />

96. Bazile, D. et al., Stealth MePEG-PLA nanoparticles avoid uptake <strong>by</strong> the mononuclear phagocyte<br />

system, J. Pharm. Sci., 84, 493, 1995.<br />

97. Peracchia, M. T. et al., Development of sterically stabilized poly(isobutyl 2-cyanoacrylate) nanoparticles<br />

<strong>by</strong> chemical coupling of poly(ethylene glycol), J. Biomed. Mater. Res., 34, 846, 1997.<br />

98. Peracchia, M. T. et al., Synthesis of novel MePEGcyanoacrylate–hexadecylcyanoacrylate amphiphilic<br />

copolymer for nanoparticle technology, Macromolecules, 30, 846, 1997.<br />

99. Peracchia, M. T. et al., Biodegradable PEGylated nanoparticles from a novel MePEGcyanoacrylate–<br />

hexadecylcyanoacrylate amphiphilic copolymer, Pharm. Res., 15, 548, 1998.<br />

100. Couvreur, P. et al., Tissue distribution of antitumor drugs associated with polyalkylcyanoacrylate<br />

nanoparticles, J. Pharm. Sci., 69, 199, 1980.<br />

101. Peracchia, M. T., Stealth w PEGylated polycyanoacrylate nanoparticles for intravenous administration<br />

and splenic targeting, J. Control. Release, 60, 121, 1999.<br />

102. Peracchia, M. T. et al., Visualization of in vitro protein rejecting properties of PEGylated stealth<br />

polycyanoacrylate nanoparticles, Biomaterials, 20, 1269, 1999.<br />

103. deGennes, P. G., Confirmation of polymers attached to an interface, Macromolecules, 13, 1069,<br />

1980.<br />

104. Peracchia, M. T. et al., Complement consumption <strong>by</strong> poly(ethylene glycol) in different confirmations<br />

chemically coupled to poly(isobutyl-2-cyanoacrylate) nanoparticles, Life Sci., 61, 749, 1997.<br />

105. Ping, L. Y. et al., PEGylated polycyanoacrylate nanoparticles as salvicine carriers: Synthesis, preparation,<br />

and in vitro characterization, Acta Pharmacol. Sin., 22, 645, 2001.<br />

106. Kante, B. et al., Toxicity of polalkylcyanoacrylate nanoparticles. I. Free nanoparticles, J. Pharm.<br />

Sci., 71, 790, 1982.<br />

107. Couvreur, P. et al., Toxicity of polyalkylcyanoacrylate nanoparticles. II. Doxorubicin-loaded nanoparticles,<br />

J. Pharm. Sci., 71, 786, 1982.<br />

108. Olivier, J. C. et al., Indirect evidence that drug brain targeting using polysorbate 80-coated polybutylcyanoacrylate<br />

nanoparticles is related to toxicity, Pharm. Res., 16, 1836, 1999.<br />

109. Simeonova, M., Chorbadjiev, K., and Antcheva, M., Study of the effect of polybutylcyanoacrylate<br />

nanoparticles and their metabolites on the primary immune response in mice to sheep red blood cells,<br />

Biomaterials, 19, 2187, 1998.<br />

110. Couvreur, P. et al., Biodegrdable polymeric nanoparticles as drug carrier for antitumor agents, In<br />

Polymeric Nanoparticles and Microspheres, Guiot, P. and Couvreur, P., Eds., CRC Press, Boca<br />

Raton, FL, pp. 27–93, 1986.<br />

111. Gelperina, S. E. et al., Toxicological studies of doxorubicin bound to polysorbate 80-coated poly<br />

(butyl cyanoacrylate) nanoparticles in healthy rats and rats with intracranial glioblastoma, Toxicol.<br />

Lett., 126, 131, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 285<br />

112. Jefferiss, C. D., Lee, A. J., and Ling, R. S., Thermal aspects of self-curing polymethylmethacrylate,<br />

J. Bone Joint Surg. Br., 57, 511, 1975.<br />

113. Vinters, H. V. et al., The histotoxicity of cyanoacrylates: A selective review, Neuroradiology, 27,<br />

279, 1985.<br />

114. Vinters, H. V., Lundie, M. J., and Kaufmann, J. C., Long-term pathological follow-up of cerebral<br />

arteriovenous malformations treated <strong>by</strong> embolization with bucrylate, N. Engl. J. Med., 314, 477,<br />

1986.<br />

115. Tranbahuy, P. et al., Direct intratumoral embolization of juvenile angiofibroma, Am. J. Otolaryngol.,<br />

15, 429, 1994.<br />

116. San Millan Ruiz, D. et al., Pathology findings with acrylic implants, Bone, 25(Suppl. 2), 85S,<br />

1999.<br />

117. Matsumoto, T., Carcinogenesis and cyanoacrylate adhesives, J. Am. Med. Assoc., 202, 1057, 1967.<br />

118. Samson, D. and Marshall, D., Carcinogenic potential of isobutyl-2-cyanoacrylate, J. Neurosurg., 65,<br />

571, 1986.<br />

119. Hatanaka, S. et al., Induction of malignant fibrous histiocytoma in female Fisher rats <strong>by</strong> implantation<br />

of cyanoacrylate, zirconia, polyvinyl chloride or silicone, In Vivo, 7, 111, 1993.<br />

120. Canter, H. I. et al., Tissue response to N-butyl-2-cyanoacrylate after percutaneous injection into<br />

cutaneous vascular lesions, Ann. Plast. Surg., 49, 520, 2002.<br />

121. Kattan, J. et al., Phase I clinical trial and pharmacokinetic evaluation of doxorubicin carried <strong>by</strong><br />

polyisohexylcyanoacrylate nanoparticles, Invest. New Drugs, 10, 191, 1992.<br />

122. Gibaud, S. et al., Increased bone marrow toxicity of doxorubicin bound to nanoparticles, Eur.<br />

J. Cancer, 30, 820, 1994.<br />

123. Chiannilkulchai, N. et al., Hepatic tissue distribution of doxorubicin-loaded nanoparticles after i.v.<br />

administration in reticulosarcoma M5076 metastasis-bearing mice, Cancer Chemother. Pharmacol.,<br />

26, 122, 1990.<br />

124. Zhang, Z. R., Liao, G. T., and Hou, S. X., Study on mitoxantrone polycyanoacrylate nanospheres,<br />

Yao Xue Xue Bao, 29, 544, 1994.<br />

125. Yang, Y. X. et al., Antitumor activity of mitoxantrone-nanosphere against murine liver tumor H22,<br />

Sichuan Da Xue Xue Bao Yi Xue Ban, 35, 68, 2004.<br />

126. Soma, C. E. et al., Investigation of the role of macrophages on the cytotoxicity of doxorubicin and<br />

doxorubicin-loaded nanoparticles on M5076 cells in vitro, J. Control. Release, 68, 283, 2000.<br />

127. Bogush, T. et al., Direct evaluation of intracellular accumulation of free and polymer-bound anthracyclines,<br />

Cancer Chemother. Pharmacol., 35, 501, 1995.<br />

128. Bennis, S. et al., Enhanced cytotoxicity of doxorubicin encapsulated in polyisohexylcyanoacrylate<br />

nanospheres against multidrug-resistant tumor cells in culture, Eur. J. Cancer, 30A, 89, 1994.<br />

129. Simeonova, M. and Antcheva, M., In vitro study of cytotoxic activity of vinblastine in a free form<br />

and associated with nanoparticles, Acta Physiol. Pharmacol. Bulg., 20, 31, 1994.<br />

130. Rusakov, I. G. et al., Use of cytostatics in polymer-drug complexes in the treatment of malignant<br />

neoplasms of the maxilla, Sov. Med., 1, 23, 1990.<br />

131. Pan, W. S. and Hu, J., Studies of 5-fluorouracil nanocapsules, Yao Xue Xue Bao, 26, 280, 1991.<br />

132. Jiang, X. H. et al., The antihepatoma effect of lyophilized aclacinomycin A polyisobutylcyanoacrylate<br />

nanoparticles in vitro and in vivo, Yao Xue Xue Bao, 30, 179, 1995.<br />

133. Brasseur, F. et al., Actinomycin D absorbed on polymethylcyanoacrylate nanoparticles: increased<br />

efficiency against an experimental tumor, Eur. J. Cancer, 16, 1441, 1980.<br />

134. Chissov, V. I. et al., Immediate results of the use of prospidin in a cyanoacrylate glue compound in<br />

cancer of the esophagus, Sov. Med., 7, 28, 1989.<br />

135. Pérez, M. et al., Use of N-butyl-2-cyanoacrylate in oral surgery: biological and clinical evaluation,<br />

Artif. Organs, 24, 241, 2000.<br />

136. Walker, J., Martin, C., and Callaghan, R., Inhibition of P-glycoprotein function <strong>by</strong> XR9576 in a solid<br />

tumor model can restore anticancer drug efficacy, Eur. J. Cancer, 40, 594, 2004.<br />

137. Cuvier, C. et al., Doxorubicin-loaded nanospheres <strong>by</strong>pass tumor cell multidrug resistance, Biochem.<br />

Pharmacol., 44, 509, 1992.<br />

138. Vauthier, C. et al., Poly(alkylcyanoacrylates) as biodegradable materials for biomedical applications,<br />

Adv. Drug Deliv. Rev., 55, 519, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


286<br />

Nanotechnology for Cancer Therapy<br />

139. Colin de Verdiere, A. et al., Uptake of doxorubicin from loaded nanoparticles in multidrug-resistant<br />

leukemic murine cells, Cancer Chemother. Pharmacol., 3, 504, 1994.<br />

140. Laurand, A. et al., Quantification of the expression of multidrug resistance-related genes in human<br />

tumor cell lines grown with free doxorubicin or doxorubicin encapsulated in polyisohexylcyanoacrylate<br />

nanospheres, Anticancer Res., 24, 3781, 2004.<br />

141. Pepin, X. et al., On the use of ion-pair chromatography to elucidate doxorubicin release mechanism<br />

from polyalkyl cyanoacrylate nanoparticles at the cellular level, J. Chromatogr. B, 702, 181, 1997.<br />

142. Hu, Y.-P. et al., On the mechanism of action of doxorubicin encapsulation in nanospheres for the<br />

reversal of multidrug resistance, Cancer Chemother. Pharmacol., 37, 556, 1996.<br />

143. Vauthier, C. et al., Drug delivery to resistant tumors: the potential of polyalkyl cyanoacrylate<br />

nanoparticles, J. Control. Release, 93, 151, 2003.<br />

144. Kobayashi, H., Takamura, Y., and Miyachi, H., Novel approaches to reversing anticancer drug<br />

resistance using gene-specific therapeutics, Hum. Cell., 14, 172, 2001.<br />

145. Juliano, R. L. et al., Antisense pharmacodynamics: critical issues in the transport and delivery of<br />

antisense oligonucleotides, Pharm. Res., 16, 494, 1999.<br />

146. Maia, R. C., Carrico, M. K., Klumb, C. E., Noronha, H., Coelho, A. M., Vasconcelos, F. C., and<br />

Ruimanek, V. M., Clinical approach to circumvention, of multidrug resistance in refractory<br />

leukaemic patients: Association of cyclosporine A with etoposide, J. Exp. Clin. Cancer Res., 16,<br />

419, 1997.<br />

147. Toffoli, G. et al., Reversal activity of cyclosporine A and its metabolites M1 M17 and M21 in<br />

multidrug resistant cells, Int. J. Cancer, 71, 900, 1997.<br />

148. Den Boer, M. L. et al., The modulating effect of PSC 833 cyclosporine A, verapamil and genistein on<br />

in vitro cytotoxicity and intracellular content of daunorubicin in childhood acute lymphoblastic<br />

leukemia, Leukemia, 12, 912, 1998.<br />

149. Soma, C. E. et al., Reversion of multidrug resistance <strong>by</strong> co-encapsulation of doxorubicin and<br />

cyclosporine A in polyalkyl cyanoacrylate nanoparticles, Biomaterials, 21, 1, 2000.<br />

150. Colin de Verdiere, A. et al., Reversion of multidrug resistance with polyalkyl cyanoacrylate nanoparticles;<br />

towards a mechanism of action, Br. J. Cancer, 76, 198, 1997.<br />

151. Soma, C. E. et al., Ability of doxorubicin-loaded nanoparticles to overcome multidrug resistance of<br />

tumor cells after their capture <strong>by</strong> macrophages, Pharm. Res., 16, 1710, 1999.<br />

152. Povarov, L. S. et al., Partial circumvention of P-glycoprotein (Pgp)-associated resistance to doxorubicin<br />

(Dox) in MCF7/R human breast carcinoma and P388/R murine leukemia cell lines <strong>by</strong><br />

doxorubicin-14-O-hemiadipate (H-Dox), Russ. J. Bioorg. Chem., 21, 797, 1995.<br />

153. Leontieva, O. V., Preobrazhenskaya, M. N., and Bernacki, R. J., Partial circumvention of P-glycoprotein-mediated<br />

multidrug resistance <strong>by</strong> doxorubicin-14-O-hemiadipate, Invest. New Drugs, 20, 35,<br />

2002.<br />

154. Krishna, R. and Mayer, L. D., Liposomal doxorubicin circumvents PSC 833-free drug interactions,<br />

resulting in effective therapy of multidrug-resistant solid tumors, Cancer Res., 57, 5246, 1997.<br />

155. Goren, D. et al., Nuclear delivery of doxorubicin via folate-targeted liposomes with <strong>by</strong>pass of<br />

multidrug-resistance efflux pumps, Clin. Cancer Res., 6, 1949, 2000.<br />

156. Carr, B. I., Hepatocellular carcinoma: Current management and future trends, Gastroenterology,<br />

127, S218, 2004.<br />

157. Talmadge, J. E. and Hart, I. R., Inhibited growth of a reticulum cell sarcoma (M5076) Induced<br />

in vitro and in vivo <strong>by</strong> macrophage-activating agents, Cancer Res., 44, 2446, 1984.<br />

158. Talmadge, J. E. et al., Immunomodulatory and immunotherapeutic properties of recombinant<br />

g-interferon and recombinant tumor necrosis factor in mice, Cancer Res., 47, 2563, 1987.<br />

159. Iigo, M., Shimizu, I., and Sagawa, K., Synergistic antitumor effects of carboplatin and interferons on<br />

hepatic metastases of colon carcinoma 26 and M5076 reticulum cell sarcoma, Jpn. J. Cancer Res.,<br />

84, 794, 1993.<br />

160. Barraud, L. et al., Increase of doxorubicin sensitivity <strong>by</strong> doxorubicin-loading into nanoparticles for<br />

hepatocellular carcinoma cells in vitro and in vivo, J. Hepatol., 42, 736, 2005.<br />

161. Zhi-Rong, Z. et al., Study on the anticarcinogenic effect and acute toxicity of liver-targeting mitoxantrone<br />

nanoparticles, World J. Gastroenterol., 5, 511, 1999.<br />

162. Tamai, I. and Tsuji, A., Drug delivery through the blood–brain barrier, Adv. Drug. Deliv. Rev., 19,<br />

401, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Poly(Alkyl Cyanoacrylate) Nanoparticles for Delivery of Anti-Cancer Drugs 287<br />

163. Vajkoczy, P. and Menger, M. D., Vascular microenvironment in gliomas, J. Neurooncol., 50, 99,<br />

2000.<br />

164. Olivier, J.-C., Drug transport to brain with targeted nanoparticles. NeuroRx (The American Society<br />

for Experimental NeuroTherapeutics, Inc.), 2, 108, 2005.<br />

165. Rautioa, J. and Chikhale, P. J., Drug delivery systems for brain tumor therapy, Curr. Pharm. Des.,<br />

10, 1341, 2004.<br />

166. Kreuter, J., Influence of the surface properties on nanoparticle-mediated transport of drugs to the<br />

brain, J. Nanosci. Nanotechnol., 4, 484, 2004.<br />

167. Calvo, P. et al., Quantification and localization of PEGylated polycyanoacrylate nanoparticles in<br />

brain and spinal cord during experimental allergic encephalomyelitis in the rat, Eur. J. Neurosci., 15,<br />

1317, 2002.<br />

168. Brigger, I. et al., Poly(ethylene glycol)-coated hexadecylcyanoacrylate nanospheres display a<br />

combined effect for brain tumor targeting, J. Pharmacol. Exp. Ther., 303, 928, 2002.<br />

169. Stouch, T. R. and Gudmundsson, O., Progress in understanding the structure–activity relationships of<br />

P-glycoprotein, Adv. Drug Deliv. Rev., 54, 315, 2002.<br />

170. Doyle, L. A. et al., A multidrug resistance transporter from human MCF-7 breast cancer cells, Proc.<br />

Natl. Acad. Sci. USA, 95, 15665, 1998.<br />

171. Aouali, N. et al., Enhanced cytotoxicity and nuclear accumulation of doxorubicin-loaded nanospheres<br />

in human breast cancer MCF7 cells expressing MRP1, Int. J. Oncol., 23, 1195, 2003.<br />

172. Henry-Toulme, N., Grouselle, M., and Ramaseilles, C., Multidrug resistance <strong>by</strong>pass in cells exposed<br />

to doxorubicin loaded nanospheres absence of endocytosis, Biochem. Pharmacol., 50, 1135, 1995.<br />

173. Marbeuf-Gueye, C. et al., Kinetics of anthracycline efflux from multidrug resistance protein-expressing<br />

cancer cell compared with P-glycoprotein-expressing cancer cells, Mol. Pharmacol., 53, 141,<br />

1998.<br />

174. Kisara, S. et al., Combined effects of buthionine sulfoximine and cepharanthine on cytotoxic activity<br />

of doxorubicin to multidrug-resistant cells, Oncol. Res., 7, 191, 1995.<br />

175. Nishioka, Y. and Yoshino, H., Lymphatic targeting with nanoparticulate system, Adv. Drug Deliv.<br />

Rev., 47, 55, 2001.<br />

176. Hashida, M., Egawa, M., Muranishi, S., and Sezaki, H., Role of intra-muscular administration of<br />

water-in-oil emulsions as a method for increasing the delivery of anticancer agents to regional<br />

lymphatics, J. Pharmacokinet. Biopharm., 5, 223, 1997.<br />

177. Parker, R. J., Priester, E. R., and Sieber, S. M., Comparison of lymphatic uptake, metabolism,<br />

excretion and biodistribution of free and liposome entrapped [ 14 C]cytosine b-D-arabinofuranoside<br />

following intraperitoneal administration to rats, Drug Metab. Dispos., 10, 40, 1982.<br />

178. Porter, C. J., Drug delivery to the lymphatic system, Crit. Rev. Ther. Drug Carrier Syst., 14, 333,<br />

1997.<br />

179. Reddy, L. H. et al., Influence of administration route on tumor uptake and biodistribution of etoposide<br />

loaded solid lipid nanoparticles in Dalton’s lymphoma tumor bearing mice, J. Control. Release,<br />

105, 185, 2005.<br />

180. Gupta, P. K. and Hung, C. T., Magnetically controlled targeted micro-carrier systems, Life Sci., 44,<br />

175, 1989.<br />

181. Widder, K. J. et al., Selective targeting of magnetic albumin microspheres containing low-dose<br />

doxorubicin: total remission in Yoshida sarcomabearing rats, Eur. J. Cancer Clin. Oncol., 19,<br />

135, 1983.<br />

182. Rudge, S. et al., Adsorption and desorption of chemotherapeutic drugs from a magnetically targeted<br />

carrier (MTC), J. Control. Release, 74, 335, 2001.<br />

183. Gao, H. et al., Preparation of magnetic polybutyl cyanoacrylate nanospheres encapsulated with<br />

aclacinomycin A and its effect on gastric tumor, World J. Gastroenterol., 10, 2010, 2004.<br />

184. Yeo, W. et al., A phase II study of doxorubicin (A) versus cisplatin (P)/interferon g-2b (I)/doxorubicin<br />

(A)/fluorouracil (F) combination chemotherapy (PIAF) for inoperable hepatocellular carcinoma<br />

(HCC), Proc. Am. Soc. Clin. Oncol., 23, 4026, 2004.<br />

185. Loewe, C. et al., Arterial embolization of unresectable hepatocellular carcinoma with use of cyanoacrylate<br />

and lipiodol, J. Vasc. Interv. Radiol., 13, 61, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


288<br />

Nanotechnology for Cancer Therapy<br />

186. Loewe, C. et al., Permanent transarterial embolization of neuroendocrine metastases of the liver<br />

using cyanoacrylate and lipiodol: Assessment of mid- and long-term results, Am. J. Roentgenol., 180,<br />

1379, 2003.<br />

187. Rand, T. et al., Arterial embolization of unresectable hepatocellular carcinoma with use of microspheres,<br />

lipiodol, and cyanoacrylate, Cardiovasc. Intervent. Radiol., 28, 313, 2005.<br />

188. Harman, M., Etlik, O., and Unal, O., Direct percutaneous embolization of a carotid body tumor with<br />

n-butyl cyanoacrylate: An alternative method to endovascular embolization, Acta Radiol., 45, 646,<br />

2004.<br />

189. Yamagami, T. et al., Value of transcatheter arterial embolization with coils and n-butyl cyanoacrylate<br />

for long-term hepatic arterial infusion chemotherapy, Radiology, 230, 792, 2004.<br />

190. Berghammer, P. et al., Arterial hepatic embolization of unresectable hepatocellular carcinoma using<br />

a cyanoacrylate/lipiodol mixture, Cardiovasc. Intervent. Radiol., 21, 214, 1998.<br />

191. Inaba, Y. et al., Right gastric artery embolization to prevent acute gastric mucosal lesions in patients<br />

undergoing repeat hepatic arterial infusion chemotherapy, J. Vasc. Interv. Radiol., 12, 957, 2001.<br />

192. Nadalini, V. F. et al., Therapeutic occlusion of the hypogastric arteries with isobutyl-2-cyanoacrylate<br />

in vesical and prostatic cancer, Radiol. Med. (Torino), 67, 61, 1981.<br />

193. Papo, J., Baratz, M., and Merimsky, E., Infarction of renal tumors using isobutyl-2 cyanoacrylate and<br />

lipiodol, AJR Am. J. Roentgenol., 137, 781, 1981.<br />

194. Carmignani, G., Belgrano, E., and Puppo, P., First clinical results with isobutyl-2-cyanoacrylate in<br />

the arterial embolisation of cancer of the kidney, Radiol. Med. (Torino), 64, 991, 1978.<br />

195. Abud, D. G. et al., Intratumoral injection of cyanoacrylate glue in head and neck paragangliomas,<br />

Am. J. Neuroradiol., 25, 1457, 2004.<br />

196. Derdeyn, C. P., Direct puncture embolization for paragangliomas: promising results but preliminary<br />

data, Am. J. Neuroradiol., 25, 1453, 2004.<br />

197. Patetsios, P. et al., Management of carotid body paragangliomas and review of a 30-year experience,<br />

Ann. Vasc. Surg., 16, 331, 2002.<br />

198. Muhm, M. et al., Diagnostic and therapeutic approaches to carotid body tumors. Review of 24<br />

patients, Arch. Surg., 132, 279, 1997.<br />

199. Marchesi, M. et al., Surgical treatment of paragangliomas of the neck, Int. Surg., 82, 394, 1997.<br />

200. Casasco, A. et al., Devascularization of craniofacial tumors <strong>by</strong> percutaneous tumor puncture, Am.<br />

J. Neuroradiol., 15, 1233, 1994.<br />

201. Pierot, L. et al., Embolization <strong>by</strong> direct puncture of hypervascularized oral tumors, Ann. Otolaryngol.<br />

Chir. Cervicofac., 111, 403, 1994.<br />

202. Schwab, G. et al., Antisense oligonucleotides adsorbed to polyalkylcyanoacrylate nanoparticles<br />

specifically inhibit mutated Ha-ras-mediated cell proliferation and tumorigenicity in nude mice,<br />

Proc. Natl. Acad. Sci. USA, 91, 10460, 1994.<br />

203. Lambert, G., Fattal, E., and Couvreur, P., Nanoparticulate systems for the delivery of antisense<br />

oligonucleotides, Adv. Drug Deliv. Rev., 47, 99, 2001.<br />

204. Chavany, C. et al., Polyalkylcyanoacrylate nanoparticles as polymeric carriers for antisense oligonucleotides,<br />

Pharm. Res., 9, 441, 1992.<br />

205. Zimmer, A., Antisense oligonucleotide delivery with poly-hexylcyanoacrylate nanoparticles as<br />

carriers, Methods: A Companion to Methods in Enzymol., 18, 286, 1999.<br />

206. Kircheis, R., Tumor targeting with surface-shielded ligand–polycation DNA complexes, J. Control.<br />

Release, 72, 165, 2001.<br />

207. Nakada, Y. et al., Pharmacokinetics and biodistribution of oligonucleotide adsorbed onto poly(isobutylcyanoacrylate)<br />

nanoparticles after intravenous administration in mice, Pharm. Res., 13, 38,<br />

1996.<br />

208. Lambert, G. et al., Effect of polyisobutylcyanoacrylate nanoparticles and lipofectin loaded with<br />

oligonucleotides on cell viability and PKC alpha neosynthesis in HepG2 cells, Biochimie, 80,<br />

969, 1998.<br />

209. Lambert, G. et al., Polyisobutylcyanoacrylate nanocapsules containing an aqueous core for the<br />

delivery of oligonucleotides, Int. J. Pharm., 214, 13, 2001.<br />

210. Schroeder, U. et al., Nanoparticle technology for delivery of drugs across the blood–brain barrier,<br />

J. Pharm. Sci., 87, 1305, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


16<br />

CONTENTS<br />

Aptamers and Cancer<br />

Nanotechnology<br />

Omid C. Farokhzad, Sangyong Jon, and Robert Langer<br />

16.1 Targeted Drug Delivery: An Introduction ........................................................................ 289<br />

16.2 Nucleic Acid Ligands (Aptamers): From Discovery to Practice ..................................... 291<br />

16.2.1 Physicochemical Properties of Aptamers ........................................................... 292<br />

16.2.2 Methods for Isolation of Aptamers..................................................................... 293<br />

16.2.3 Examples of Aptamers for Targeted Delivery ................................................... 294<br />

16.2.3.1 Alpha-v Beta-3 ................................................................................... 294<br />

16.2.3.2 Human Epidermal Growth Factor-3 .................................................. 295<br />

16.2.3.3 Prostate Specific Membrane Antigen ................................................ 295<br />

16.2.3.4 Nucleolin ............................................................................................ 295<br />

16.2.3.5 Sialyl Lewis X.................................................................................... 295<br />

16.2.3.6 Cytotoxic T-Cell Antigen-4 ............................................................... 296<br />

16.2.3.7 Fibrinogen-Like Domain of Tenascin-C............................................ 296<br />

16.2.3.8 Pigpen ................................................................................................. 296<br />

16.3 Aptamer–Nanoparticle Conjugates for Targeted Cancer Therapy................................... 296<br />

16.3.1 Properties of Nanoparticles for Conjugation to Aptamers ................................. 297<br />

16.3.1.1 Size of Nanoparticles ......................................................................... 297<br />

16.3.1.2 Polymers for Synthesis of Nanoparticles........................................... 298<br />

16.3.1.3 Charge of Nanoparticles..................................................................... 298<br />

16.3.1.4 Surface Modification of Nanoparticles .............................................. 299<br />

16.3.2 Conjugation Strategies for Nanoparticle–Aptamer Conjugates ......................... 300<br />

16.4 Aptamer–Drug Conjugates for Targeted Cancer Therapy ............................................... 301<br />

16.5 Aptamer–Nanoparticle Conjugates for Protein Detection................................................ 304<br />

16.5.1 Aptamer–QD Conjugates for the Detection of Proteins .................................... 304<br />

16.5.2 Aptamer–Gold-Nanoparticle Conjugates for the Detection of Proteins ............ 305<br />

16.6 Conclusion ......................................................................................................................... 306<br />

Acknowledgments ........................................................................................................................ 306<br />

References..................................................................................................................................... 306<br />

16.1 TARGETED DRUG DELIVERY: AN INTRODUCTION<br />

With advances in nanotechnology, it is becoming increasingly possible to combine specialized<br />

delivery vehicles and targeting approaches to develop highly selective and effective cancer<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

289


290<br />

Nanotechnology for Cancer Therapy<br />

therapeutic and diagnostic modalities. 1–4 In the case of therapeutic vehicles, drug targeting is<br />

expected to reduce the undesirable and sometimes life-threatening side effects common in anticancer<br />

therapy. Drug targeting can also enhance the efficacy of drugs because the drug is delivered<br />

directly to cancer cells, making it possible to deliver a cytotoxic dose of the drug in a<br />

controlled manner.<br />

Cancer targeting may be achieved passively <strong>by</strong> continuously concentrating drug encapsulated<br />

nanoparticles in the tumor interstitial space due to the enhanced permeability and retention (EPR)<br />

effect. 5,6 This process occurs because of a differentially large quantity of nanoparticles extravasting<br />

out of tumor microvasculature, leading to an accumulation of drugs in the tumor interstitium.<br />

Multiple factors influence the EPR, including active angiogenesis and high vascularity, defective<br />

vascular architecture, impaired lymphatic drainage, and extensive production of vascular<br />

mediators, such as bradykinin, nitric oxide, vascular endothelial growth factor (VEGF), prostaglandins,<br />

collagenase, and peroxynitrite. The ease of vehicle extravasation is a function of the<br />

maximum size of the transvascular transport pathways in tumor microvasculature and is determined<br />

mainly through the size of open interendothelial gap junctions and trans-endothelial channels. The<br />

pore cutoff size of these transport pathways has been estimated between 400 and 600 nm, and<br />

extravasation of liposomes into tumors in vivo suggests a cutoff size in the range of 400 nm. 8 As a<br />

general rule, particle extravasation is inversely proportional to size, and small particles (fewer than<br />

200 nm size) should be most effective for extravasating the tumor microvasculature. 8–10 Passive<br />

tumor targeting has several limitations, including the inability to achieve a sufficiently high level of<br />

drug concentration at the tumor site, resulting in lack of drug efficacy. 11 Additionally, the lack of<br />

targeting efficiency contributes to undesirable systemic adverse effects (for reviews, see). 4,12<br />

Active tumor targeting may be achieved <strong>by</strong> both local and systemic administration of specially<br />

designed vehicles that recognize biophysical characteristics that are unique to the cancer cells.<br />

Most commonly, this represents binding of vehicles to target antigens using ligands that recognize<br />

tumor-specific or tumor-associated antigens. Some of these therapeutic conjugates are now under<br />

clinical development or are in clinical practice today. In the case of local drug delivery, such as<br />

through the injection of delivery vehicles within an organ, it is possible to achieve a desired effect<br />

within a subset of cells, as opposed to a generalized effect on all the cells of the targeted organ.<br />

In the case of cancer, the cytotoxic effects of a therapeutic agent would be directed to cancer cells<br />

while minimizing harm to noncancerous cells within and outside of the targeted organ. For<br />

example, suicide-targeted gene delivery has been demonstrated to be effective in killing prostate<br />

cancer but not healthy muscle cells in xenograft mouse models of prostate cancer. 13 Similar<br />

approaches have been used to deliver docetaxel via intratumoral injection of targeted controlled<br />

release polymer vehicles to prostate cancer cells in nude mice. 14 This approach is particularly<br />

useful for primary tumors that have not yet metastasized, such as localized breast or<br />

prostate cancer.<br />

For metastatic cancers, the drug delivery vehicle would ideally be administered systemically<br />

because the location, abundance, and size of tumor metastasis within the body limits its visualization<br />

or accessibility, thus making local delivery approaches impractical. Despite the obvious<br />

advantages of systemic delivery, this approach is more challenging, and many physiological and<br />

biochemical barriers must be overcome for vehicles to reach the tumor and ultimately be capable of<br />

delivering therapeutically effective concentrations of the drugs directly to the cancer cells. One<br />

major challenge is the early systemic clearance of vehicles <strong>by</strong> the mononuclear phagocytic cells<br />

present in the liver, spleen, lung, and bone marrow. 15–18 This is in part due to the large percentage of<br />

cardiac output directed to these organs, and in part due to the dense population of macrophages and<br />

monocytes that home in these organs. Additionally, a variety of factors play an important role in the<br />

ultimate success of systemic targeted approaches, including those factors related to the<br />

biochemical and physical characteristics of the nanoparticles such as: (1) the chemical properties<br />

of the controlled release polymer system and the encapsulated drug; (2) the size of the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 291<br />

nanoparticles; (3) surface charge and surface hydrophilicity of nanoparticles; and (4) the chemical<br />

nature of the targeting molecules. 19<br />

Several classes of molecules have been utilized for targeting applications, including various<br />

forms of antibody-based molecules, 20 such as chimeric human–murine antibodies, humanized<br />

antibodies, single chain Fv generated from murine hybridoma or phage display, and minibodies. 58<br />

Multivalent antibody-based targeting structures, such as multivalent minibodies, single-chain<br />

dimers and dibodies, and multispecific binding proteins, including bispecific antibodies and<br />

antibody-based fusion proteins, have all been evaluated. More recently, other classes of ligands,<br />

such as carbohydrates 21 and nucleic acid ligands, also called aptamers, 22,23 have been used as escort<br />

molecules for targeted delivery applications. The concept of nucleic acid molecules acting as<br />

ligands was first described in the 1980s when it was shown that some viruses encode small<br />

structured RNA that binds to viral and host proteins with high affinity and specificity. These<br />

RNA nucleic acid ligands had evolved over time to enhance the survival and propagation of the<br />

viruses. Subsequently, it was shown that these naturally occurring RNA ligands can inhibit the viral<br />

replication and have therapeutic benefits. 24 Finally, methods were developed to perform in vitro<br />

evolution and to isolate novel nucleic acid ligands that bind to a myriad of important molecules for<br />

diverse applications in research and clinical practice. 25,26<br />

16.2 NUCLEIC ACID LIGANDS (APTAMERS): FROM<br />

DISCOVERY TO PRACTICE<br />

The feasibility of using antibodies for targeted therapy, particularly for oncologic diseases, has been<br />

demonstrated repeatedly in the literature. Rituximab (Rituxane) was the first therapeutic based on<br />

monocloncal antibodies to receive FDA approval in 1997 for treating patients with relapsed or<br />

refractory low-grade or follicular, CD20 positive, B-cell non-Hodgkin’s lymphoma. 27 Awide<br />

variety of antibody-based drugs are now under clinical development or are in clinical practice<br />

today. For example, denileukin diftitox (Ontake) is an FDA-approved immunotoxin for the<br />

treatment of cutaneous T-cell lymphoma. 28 Many other radioimmunoconjugates or chemoimmunoconjugates<br />

directed against cell surface antigens are currently in various stages of clinical and<br />

pre-clinical development. Despite the recent success of monoclonal antibodies as targeting<br />

moieties, the use of antibodies for drug targeting may have a number of potential disadvantages.<br />

First, antibodies are large molecules (w20 nm for intact antibodies) 29,30 and their use in developing<br />

nanoscale therapeutic and diagnostic tools may result in an increase in vehicle size without added<br />

advantage. Second, antibodies may be immunogenic. Despite the current engineering approaches to<br />

yield improved humanized antibodies, this problem is not universally solved. Third, the biological<br />

development of monoclonal antibodies can be difficult and unpredictable. For example, the target<br />

antigen may not be well tolerated <strong>by</strong> the animal used to produce the antibodies, or the target<br />

molecules may be inherently less immunogenic, making it difficult to raise antibodies against<br />

such targets (although, this problem is overcome with the use of phage display libraries). 31,32<br />

Fourth, the production of antibodies involves a biological process that can result in batchto-batch<br />

variability in their performance, particularly when production is scaled up. The ideal<br />

targeting molecule for the delivery of nanoscale therapeutic and diagnostic systems should, like<br />

monoclonal antibodies, bind with high affinity and specificity to a target antigen but overcome or<br />

ameliorate some of the problems associated with the use and production of monoclonal antibodies.<br />

Nucleic acid aptamers are a novel class of ligands 25,26 that are small, nonimmunogenic, easy to<br />

isolate, characterize and modify, and exhibit high specificity and affinity for their target antigen.<br />

Aptamers derive their name from the Latin word aptus, meaning “to fit.” In the short time since Jack<br />

Szostak and Larry Gold independently described the groundbreaking methodology for in vitro<br />

evolution of aptamers, this class of ligands has emerged as an important arsenal in research and<br />

clinical medicine. 22,33 Considering the many favorable characteristics of aptamers, which have<br />

resulted in their rapid progress into clinical practice, 33 researchers have begun to exploit this class<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


292<br />

of molecules for targeted delivery of controlled release polymer drug delivery vehicles. Recently,<br />

we described the first proof-of-concept drug delivery vehicles utilizing aptamers for targeted<br />

delivery of drug encapsulated nanoparticles 23 and have gone on to the show efficacy of similarly<br />

designed nanoparticles against prostate cancer tumors in vivo (Farokhzad et al., submitted). 14<br />

16.2.1 PHYSICOCHEMICAL PROPERTIES OF APTAMERS<br />

Nanotechnology for Cancer Therapy<br />

Nucleic acid aptamers are single-stranded DNA, RNA, or unnatural oligonucleotides that fold into<br />

unique structures capable of binding to specific targets with high affinity and specificity. Aptamers<br />

are selected in vitro from a pool of (w10 14 –10 15 ) random oligonucleotides. 34 These molecules have<br />

a molecular weight in the 10–20 kD range, which is one order of magnitude lower than that of<br />

antibodies (150 kD). 35 The small size of aptamers (w5 nm for 30–60 base pair of aptamer) 36 is<br />

desirable when utilizing them as targeting molecules for the delivery of nanoscale drug delivery<br />

vehicles, as they do not substantially increase the overall size of the vehicle. When compared to<br />

small molecule ligands, aptamers have a larger surface area, offering more points of chemical<br />

contact with their targets. Aptamers fold through intramolecular interaction into tertiary conformations<br />

with specific binding pockets. These molecules bind with high specificity and affinity to a<br />

variety of target antigens with an equilibrium dissociation constant (K d) in the 10 pM–10 mM<br />

range. 37 Unlike antibodies, their synthesis and large scale production is an entirely chemical<br />

process that does not rely on biological systems. This can translate into lower manufacturing<br />

cost and decrease batch-to-batch variability during production, which is a significant advantage<br />

for the commercialization of this class of molecules. Furthermore, due to their small size, aptamers<br />

exhibit superior tissue penetration, 22 and because of their similarity to endogenous molecules,<br />

aptamers are believed to be less immunogenic than antibodies. 38<br />

Unlike anti-sense oligonucleotides, which are single-stranded nucleic acids that require cellular<br />

uptake and affect the synthesis of a targeted protein <strong>by</strong> hybridizing to the mRNAs that encodes it,<br />

aptamers may inhibit a protein’s function <strong>by</strong> directly binding to their targets, which may include<br />

other nucleic acids, proteins, peptides, and small molecules. Aptamers can be described <strong>by</strong> a<br />

sequence of approximately 15–60 nucleotides (A, U, T, C, and G). The conformation of the<br />

aptamer confers specificity for a target molecule <strong>by</strong> interacting with multiple domains or<br />

binding pockets. Small changes in the target molecule can foil interactions, and thus aptamers<br />

can distinguish between closely related but nonidentical targets. For example, specific RNAs were<br />

identified that have a high affinity for the bronchodilator theophylline (1,3-dimethylxanthine) yet<br />

exhibit a more than 10,000 times weaker binding affinity to caffeine (1,3,7-trimethylxanthine),<br />

which differs form theophylline only <strong>by</strong> the substitution of a methyl group at the nitrogen atom<br />

N7 position. 39 Based on their unique molecular recognition properties, aptamers have found great<br />

utility for applications in areas such as in vitro and in vivo diagnostics, analytical techniques,<br />

imaging, and therapeutics. 35,40,41<br />

Aptamers are highly stable and may tolerate a wide range of temperature, pH (w4–9), and<br />

organic solvents without loss of activity, but these molecules are susceptible to nuclease<br />

degradation and renal clearance in vivo. Therefore, their pharmacokinetic properties must be<br />

enhanced prior to in vivo applications. Several approaches have been adopted to optimize the<br />

properties of aptamers, such as: (1) capping their terminal ends; (2) substituting naturally occurring<br />

nucleotides with unnatural nucleotides that are poor substrates for nuclease degradation (i.e., 2 0 -F,<br />

2 0 -OCH3, or2 0 -NH2-modified nucleotides); (3) substituting naturally occurring nucleotides with<br />

hydrocarbon linkers; and (4) use of L-enantiomers of nucleotides to generate mirror image aptamers<br />

commonly referred to as spiegelmers. 42–45 Aptamers can also be stabilized using locked nucleic<br />

acid modifications to reduce conformational flexibility, 46 circularized, linked together in pairs, or<br />

clustered onto a substrate. Alternatively, a nuclease resistant aptamer may be selected de novo<br />

using a pool of oligonucleotides with 2 0 -F- or 2 0 -OCH 3-modified nucleotides. By combining some<br />

of these strategies, an aptamer’s half-life can be prolonged from several minutes to many hours. 35<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 293<br />

To prolong the rate of clearance of aptamers, their size may be increased <strong>by</strong> conjugation with<br />

polymers such as polyethylene glycol (PEG). 47<br />

16.2.2 METHODS FOR ISOLATION OF APTAMERS<br />

In vitro selection, 25 also called systematic evolution of ligands <strong>by</strong> exponential enrichment<br />

(SELEX), 26,48 is a protocol to isolate rare functional oligonucleotides (i.e., aptamers) from a<br />

pool of random oligonucleotides (Figure 16.1). Similar to phage display or other strategies used<br />

to isolate ligands from random libraries, SELEX is essentially an iterative selection and amplification<br />

protocol to isolate single-stranded nucleic acid ligands that bind to their target with high<br />

affinity and specificity. The complexity of the starting library is determined in part <strong>by</strong> the number of<br />

random nucleotides in the pool. For example, <strong>by</strong> using a library with 40 random nucleotides, a pool<br />

of 10 24 distinct nucleotides can be generated. Practically speaking, the number of ligands in the<br />

starting pool for in vitro selection is closer to 10 15 , representing 1 nmol of the library.<br />

In the initial step, a library of random nucleotides flanked <strong>by</strong> fixed nucleotides is generated <strong>by</strong><br />

solid phase oligonucleotide synthesis. The random nucleotides serve to add complexity to the pool<br />

while the fixed sequences are utilized for polymerase chain reaction (PCR) amplification. In the<br />

(10 14 -10 15 oligonucleotides)<br />

Nucleic acid library<br />

Fixed Random Fixed<br />

Primer 1<br />

ACGGACTACGGTTGAG-40N-CGAATGCGAACGTACAGT<br />

Primer 2<br />

Repeat until maximum<br />

enrichment of ligands<br />

has been obtained<br />

Isolation of bound<br />

nucleotides<br />

Target molecule bound<br />

to a solid support<br />

Cloning of oligonucleotides<br />

and amplification<br />

of plasmids<br />

Isolation of individual<br />

aptamers and<br />

characterization<br />

Reiterative rounds<br />

of selection and<br />

amplification<br />

FIGURE 16.1 Schematic representation of SELEX. An oligonucleotide library is synthesized containing<br />

random sequences that are flanked <strong>by</strong> fixed sequences that facilitate PCR amplification. Target molecules<br />

are incubated with this pool of oligonucleotides, and bound and unbound oligonucleotides are partitioned.<br />

Bound oligonucleotides are isolated, and iterative rounds of selection and amplification are performed with<br />

increased stringency to isolate aptamers with high specificity and affinity for the target molecule. Oligonucleotide<br />

ligands representing the aptamers are subsequently cloned in plasmids, amplified, and sequenced. The net<br />

result of this enrichment process is a small number of highly specific aptamers that are isolated from a large<br />

library of random oligonucleotides.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


294<br />

case of isolating DNA aptamers, the oligonucleotide pool is incubated with the target of interest,<br />

and the bound fragments are partitioned and directly amplified <strong>by</strong> PCR system. The resulting pool is<br />

used in a follow-up round of selection and amplification, and the process is repeated until the<br />

affinity for the target antigen plateaus. Typically, this will be achieved in six to ten rounds of<br />

SELEX. After the last round of SELEX, aptamers are cloned in plasmids, amplified, sequenced, and<br />

their binding constants are determined (Figure 16.1). These aptamers may be subject to additional<br />

modification such as size minimization to truncate the nucleotides not necessary for binding<br />

characteristics and nuclease stabilization <strong>by</strong> replacing naturally occurring nucleotides with<br />

modified nucleotides (i.e., 2 0 -F pyrimidines, 2 0 -OCH3 nucleotides) that are poor substrates for<br />

endo- and exonuclease degradation.<br />

In contrast to the isolation of DNA aptamers, which require single-step amplification after<br />

portioning, the selection of RNA aptamers involves additional steps, including transcription of<br />

an RNA pool from the starting DNA library, reverse transcription of the partitioned RNA pool to<br />

generate a cDNA fragment and subsequent amplification of DNA, and transcription into RNA for<br />

the next round of selection. 34 The advantage of RNA SELEX, however, is that unnatural nucleotides<br />

such as 2 0 -F pyrimidines and 2 0 -OCH3 nucleotides may be used in the transcription of the<br />

RNA pool because these modified bases are utilized <strong>by</strong> RNA polymerase as substrate. Furthermore,<br />

mutant RNA polymerases have also been described that are capable of improved incorporation of<br />

modified bases during transcription. 49 The resulting modified RNA pool can be used for isolation of<br />

nuclease-stable RNA aptamers. Recently, a fully 2 0 -OCH 3-modified anti-VEGF aptamer was<br />

selected, and when conjugated to 40 kDa PEG, it demonstrated a circulating half-life of 23 h. 50<br />

Conversely, a DNA polymerase that can incorporate unnatural bases such as 2 0 -F and 2 0 -OCH3 has<br />

not been described, and, consequently, DNA aptamers must be nuclease stabilized after the<br />

SELEX procedure.<br />

16.2.3 EXAMPLES OF APTAMERS FOR TARGETED DELIVERY<br />

Since their original description in 1990, aptamers have been isolated to a wide variety of targets,<br />

including intracellular proteins, transmembrane proteins, soluble proteins, carbohydrates, and small<br />

molecule drugs. As of the submission of this chapter, 624 articles related to aptamers have been<br />

published, and more than 200 aptamers have been isolated (comprehensively listed in the Aptamer<br />

Database), 51 and one aptamer, macugen (pegaptanib sodium), against the VEGF165 protein was<br />

recently approved <strong>by</strong> the FDA in December, 2004 for the treatment of neovascular macular<br />

degeneration, 52,53 underscoring the rapid progress of aptamers from its original conception to<br />

clinical application. In choosing aptamers for targeting cancer cells, the aptamer must be directed<br />

towards receptors that are preferentially or exclusively expressed on the plasma membrane of<br />

cancer cells. Alternatively, they may be delivered to extracellular matrix molecules that are<br />

expressed preferentially in tumors. To date, several tumor-specific or tumor-associated aptamers<br />

have been isolated (reviewed <strong>by</strong> Pestourie et al.). 54 A partial listing of these aptamers, as well as<br />

their size and binding characteristics, are outlined below.<br />

16.2.3.1 Alpha-v Beta-3<br />

Nanotechnology for Cancer Therapy<br />

Alpha-v beta-3 (avb3) integrin is a transmembrane glycoprotein that mediates numerous processes,<br />

including cell migration, cell growth, tumor growth and metastasis, angiogenesis, and vascular<br />

healing. The expression of this protein on the endothelial cells of tumor neovasculature is dramatically<br />

increased, making avb3 an interesting protein for targeting approaches. 55 An 85-base-pair<br />

2 0 -flouropyrimidine RNA aptamer (Apt-avb3) has been shown to bind to the avb3 protein on the<br />

surface of endothelial cells in vitro and inhibit endothelial cell proliferation and survival. 56<br />

The future in vivo use of the Apt-avb3 may require post-SELEX optimization, including size<br />

minimization to facilitate large-scale synthesis, and enhance the pharmacokinetic properties of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 295<br />

this molecule. This aptamer may be utilized for targeted cancer diagnostic and therapeutic<br />

applications.<br />

16.2.3.2 Human Epidermal Growth Factor-3<br />

Human epidermal growth factor-3 (HER-3) is a receptor tyrosine kinase that is over-expressed in<br />

several cancers. Over-expression of HER-3 is also associated with drug resistance in many HER-2<br />

over-expressed tumors, making HER-3 a candidate target for drug delivery. A30 is an RNA<br />

aptamer that was isolated against the extracellular domain of the HER-3 and is shown to inhibit<br />

heregluin dependent tyrosine phosphorylation of HER-2 and heregluin-induced growth response of<br />

MCF-7 cells at KiZ10 and 1 nM, respectively. 57 The future use of A30 for in vivo application may<br />

require post-SELEX optimization of this aptamer, including nuclease stabilization and<br />

size minimization.<br />

16.2.3.3 Prostate Specific Membrane Antigen<br />

Prostate specific membrane antigen (PSMA) encodes a folate carboxypeptidase. Its expression is<br />

tightly restricted to prostate acinar epithelium and is increased in prostatic intraepithelial neoplasia,<br />

prostatic adenocarcinoma, and in tumor-associated neovasculature. Two 2 0 -flouropyrimidine RNA<br />

aptamers, named xPSM-A9 and xPSM-A10, against the extracellular domain of the PSMA have<br />

been isolated and characterized. 58 The aptamer xPSM-A9 inhibits the enzymatic function of the<br />

PSMA noncompetitively with a K iZ2.1 nmol, and aptamer xPSM-A10 inhibits the enzymatic<br />

function of PSMA competitively with a KiZ11.9 nmol. Aptamer xPSM-A10 has also been truncated<br />

from 71 nucleotides to its current size of 56 nucleotides (w18kD).<br />

16.2.3.4 Nucleolin<br />

Nucleolin was originally described as a nuclear and cytoplasmic protein; however, a number of<br />

recent studies have shown that it can also be expressed at the cell surface. 59,60 Nucleolin is involved<br />

in the organization of the nuclear chromatin, rDNA transcription, packaging of the pre-RNA,<br />

ribosome assembly, nucleocytoplasmic transport, cytokinesis, nucleogenesis, and apoptosis.<br />

AS-1411 (formerly AGRO100) is an aptamer capable of making G-quadruplexes that bind to<br />

nucleolin on the cell surface 61 and interact with the nuclear factor kappa B (NFkB) essential<br />

modulator (NEMO) inside the cell. 62 The cytosolic localization of AS-1411 after binding to cell<br />

surface nucleolin may be exploited for the intracellular delivery of nanoparticles to cancer cells.<br />

The use of AS-1411 as a therapeutic modality has also shown promise for the treatment of cancer in<br />

humans, and Antisoma of United Kingdom is evaluating this aptamer in phase-I clinical trials. 63<br />

The therapeutic benefit of AS-1411 is presumably attributed to the disruption of the NFkB signaling<br />

inside the cells. 64<br />

16.2.3.5 Sialyl Lewis X<br />

Sialyl Lewis X (sLe x ) is a tetra saccharide glycoconjugate of transmembrane proteins that acts<br />

as a ligand for the selectin proteins during cell adhesion and inflammation. The sLe x is also<br />

abnormally overexpressed on the surface of cancer cells and may play a role in cancer cell<br />

metastasis. An RNA aptamer referred to as clone 5 has been isolated and shown to bind to the<br />

sLe x with sub-nanomolar affinity and block the sLe x /selectin mediated adhesion of HL60<br />

monocytic cell line in vitro. 65 The future use of clone 5 for in vivo targeted delivery applications<br />

may require post-SELEX optimization of this aptamer, including nuclease stabilization<br />

and size minimization.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


296<br />

16.2.3.6 Cytotoxic T-Cell Antigen-4<br />

Cytotoxic T-cell antigen-4 (CTLA-4) is a transmembrane protein that is expressed on the surface of<br />

activated T-cells and functions to attenuate the T-cell response <strong>by</strong> raising the threshold needed for<br />

T-cell activation. D60 (KdZ33 nM) is a size-minimized, 35-base-pair, 2 0 -fluoropyrimidine<br />

modified nuclease stable RNA aptamer that was selected against CTLA-4 and shown to inhibit<br />

CTLA-4 function in vitro and enhance tumor immunity in mice. 66<br />

16.2.3.7 Fibrinogen-Like Domain of Tenascin-C<br />

Tenascin-C is an extracellular matrix protein that is over-expressed during tissue remodeling<br />

processes, such as fetal development, and wound healing, as well as tumor growth. TTA1 is an<br />

aptamer (K dZ5 nM) that was isolated against the fibrinogen-like domain of Tenascin-C, 67 and may<br />

potentially be useful for cancer diagnostic and targeted therapeutic applications.<br />

16.2.3.8 Pigpen<br />

Pigpen is an endothelial protein of the Ewing’s sarcoma family that parallels the transition from<br />

quiescent to angiogenic phenotypes in vitro. Using YPEN-1 endothelial cells and N9 micorglial<br />

cells, respectively, in a selection and counter-selection strategy in SELEX, a DNA aptamer named<br />

III.1 was isolated that preferentially bound to YPEN-1 cells. 68 The III.1 was also shown to selectively<br />

bind to the microvessels of experimental rat glioblastoma. The isolation and characterization<br />

of the III.1 target identified pigpen as the target antigen. The use of III.1 aptamer for targeting the<br />

microvasculature of tumors is a potentially powerful means of delivering drugs to the site of<br />

the cancer.<br />

16.3 APTAMER–NANOPARTICLE CONJUGATES FOR TARGETED<br />

CANCER THERAPY<br />

Nanotechnology for Cancer Therapy<br />

Virtually every branch of medicine has been dramatically impacted <strong>by</strong> controlled drug delivery<br />

strategies during the past four decades, 3,69–72 including cardiology, 73 ophthalmology, 74 endocrinology,<br />

75 oncology, 76 immunology, 77 and orthopedics. 78 Drugs can be released in a controlled<br />

manner from within a material through surface or bulk erosion of the material, diffusion of the drug<br />

from the interior of the material, or swelling followed <strong>by</strong> diffusion or triggered <strong>by</strong> the environment<br />

or other external events, 2 such as changes in pH, 79 light, 80 temperature, 81 or the presence of an<br />

analyte, such as glucose. 82 In general, controlled-release polymer systems deliver drugs in the<br />

optimum dosage for long periods, thus increasing the efficacy of the drug, maximizing patient<br />

compliance and enhancing the ability to use highly toxic, poorly soluble, or relatively<br />

unstable drugs.<br />

The conjugation of aptamers to drug encapsulated nanoparticles results in targeted delivery<br />

vehicles for therapeutic application. These may include delivery of small molecule drugs, protein<br />

based drugs, nucleic acid drugs (anti-sense oligoneucleotide, RNAi or gene therapy), and targeted<br />

delivery of agents for neutron capture therapy or photodynamic therapy. Aptamers may also<br />

be bound to imaging agents to facilitate diagnosis and identification of tumor metastases. For<br />

example, it may be useful to bind aptamers to optical imaging agents including fluorophores 68<br />

and quantum dots (nanocrystals) 83 or MRI imaging agents such as magnetic nanoparticles 84,85 for<br />

detection of small foci of cancer metastasis. Multiplex systems comprising drug laden nanoparticle<br />

aptamer conjugates, together with imaging agents, represent a prospective avenue to<br />

future research.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 297<br />

16.3.1 PROPERTIES OF NANOPARTICLES FOR CONJUGATION TO APTAMERS<br />

Nanoparticles are a particularly attractive drug delivery vehicle for cancer therapeutics because<br />

they can be synthesized to recognize tumor-specific antigens and deliver drugs in a controlled<br />

manner. 4,23 The design of targeted drug delivery nanoparticles combines drug encapsulated<br />

materials, such as biodegradable polymers, with a targeting moiety (Figure 16.2). Ideally, biodegradable<br />

targeted nanoparticles should be designed with the following parameters: (1) small size<br />

(preferably between 10 and 200 nm); (2) high drug loading and entrapment efficiency; (3) low rate<br />

of aggregation; (4) slow rate of clearance from the bloodstream; and (5) optimized targeting to the<br />

desired tissue with minimized uptake <strong>by</strong> other tissues. The following sections will discuss the<br />

various parameters that must be considered for engineering of nanoparticles for targeted drug<br />

delivery applications, including the development of nanoparticle–aptamer bioconjugates.<br />

This will include discussion of nanoparticle biomaterial, size, charge, and surface modification<br />

schemes to achieve the desired design parameters. It is important to note that a detailed review is<br />

beyond the scope of this chapter, and the reader is referred to the following reviews for further<br />

information: see references. 86–90<br />

16.3.1.1 Size of Nanoparticles<br />

Biodegradable polymer<br />

Chemotherapeutic<br />

Surface functionality<br />

Spacer group<br />

Targeting moiety<br />

FIGURE 16.2 Schematic representation of targeted drug delivery vehicle composed of polymeric nanoparticles<br />

that are surface-modified with targeting agents.<br />

The biodistribution pattern of nanoparticles, active nanoparticle targeting to tumor antigens, and<br />

passive nanoparticle targeting <strong>by</strong> EPR 91 are all greatly effected <strong>by</strong> the size of the nanoparticle.<br />

Passive targeting of the nanoparticle occurs because microvasculatures of tumors are more “leaky,”<br />

thus permitting selective permeation of nanoparticles into the desired tumor tissue. This phenomenon<br />

has been exploited to target liposomes, therapeutic and diagnostic nanoparticles, and drugpolymer<br />

conjugates to cancer tissue (reviewed <strong>by</strong> Maeda). 91 Smaller particles (fewer than 150 nm)<br />

are better suited for permeating through the leaky microvasculature of the tumor cells, and their<br />

more pronounced surface curvature may also reduce interaction with the receptors on the surface of<br />

macrophages and subsequent clearance of the particles. 12 Biodistribution studies using liposomes<br />

have shown that although particles that are larger than 200 nm are largely taken up <strong>by</strong> the spleen,<br />

those less than 70 nm are also efficiently cleared <strong>by</strong> the liver. 92 Taken together, the optimal<br />

nanoparticle size should be experimentally determined for each formulation because the interplay<br />

of various parameters (polymer system, encapsulated drug, surface charge, surface modification)<br />

makes it difficult to extrapolate the ideal nanoparticle size from seemingly similar studies. In our<br />

laboratory, we have tuned the size of polymeric nanoparticles made either <strong>by</strong> emulsions or nanoprecipitation<br />

to study the best formulation for a specific application. Generally, one can make use of<br />

the interaction of different solvents used for making the emulsions or tune the size of the nanoparticle<br />

<strong>by</strong> adjusting solvent ratios and polymer concentrations in solution. Our biodistribution<br />

studies using various sizes of poly(lactic-co-glycolic acid) (PLGA)–PEG nanoparticle–aptamer<br />

bioconjugates, has suggested a linear relationship with regards to uptake <strong>by</strong> liver and spleen<br />

such that smaller particles (w80 nm) are modestly better at avoiding uptake <strong>by</strong> these organs<br />

(unpublished results).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


298<br />

16.3.1.2 Polymers for Synthesis of Nanoparticles<br />

Controlled release biodegradable nanoparticles for clinical applications can be made from a wide<br />

variety of polymers, including poly(lactic acid) (PLA), 93 poly(glycolic acid) (PGA), PLGA, 94<br />

poly(orthoesters), 95 poly(caprolactone), 96 poly(butyl cyanoacrylate), 97 polyanhydrides, 98 and<br />

poly-N-isopropylacrylamide. 99 Although many fabrication methods exist, polymeric nanoparticles<br />

are frequently made using an oil-in-water emulsion, 18,100 which involves dissolving a polymer and<br />

drug in an organic solvent, such as methylene chloride, ethyl acetate, or acetone. The organic phase<br />

is mixed with an aqueous phase <strong>by</strong> vortexing and sonicating and then evaporated, forcing the<br />

polymer to precipitate as nanoparticles in the aqueous phase. The particles are then recovered <strong>by</strong><br />

centrifugation and lyophilization.<br />

One of the considerations with respect to the material used for drug delivery is its ability to<br />

encapsulate drugs as well as degrade over the appropriate times. This subject has been an active<br />

area of investigation <strong>by</strong> our group and other investigators in academic and industry laboratories for<br />

several decades. The result has been an increasing arsenal of polymers with distinct encapsulation<br />

and release characteristics for a myriad of research, industrial, and clinical applications. 101,102 PGA<br />

and PLA are common biocompatible polymers that are used for many biomedical applications.<br />

PGA is hydrophilic because it lacks a methyl group and is more susceptible to hydrolysis, making<br />

this polymer easily degradable. Alternatively, PLA is relatively more stable in the body. 103<br />

Through these unique properties, polymers such as PLGA have been derived that are made from<br />

both glycolic acid and lactic acid components. The ability to change the ratio of these two components<br />

of the polymer can then be used to dramatically alter the rate of degradation. Therefore, <strong>by</strong><br />

choosing the desired polymer system for the synthesis of nanoparticles the rate of degradation and<br />

subsequent release of the molecule may be tuned for the intended application.<br />

16.3.1.3 Charge of Nanoparticles<br />

Nanotechnology for Cancer Therapy<br />

Nanoparticle charge has been shown to be important for regulating its pharmacokinetic properties.<br />

For example, it has been shown that anionic and cationic liposomes activate the complement system<br />

through distinct pathways, suggesting that particle charge may impact particle opsonization and<br />

phagocytosis. 104 Cationic charge on liposomes has also been shown to reduce their circulating halflife<br />

in blood and to affect their biodistribution between the tumor microvasculature and interstitium<br />

without impacting overall tumor uptake. 105 Nanoparticles could be synthesized with charged<br />

surfaces either <strong>by</strong> using charged polymers, such as poly-L-lysine, polyethylenimine (PEI), or<br />

polysaccharides, or through surface modification approaches. For example, the layer-<strong>by</strong>-layer<br />

deposition of ionic polymers have been used to change surface properties of nanoparticles, such<br />

as quantum dots, <strong>by</strong> depositing ionic polymers of interest on the charged nanoparticle surfaces. 106<br />

Furthermore, surface charge of nanoparticles has been shown to regulate their biodistribution. 107<br />

For example, increasing the charge of cationic pegylated liposomes decreases their accumulation in<br />

the spleen and blood while increasing their uptake <strong>by</strong> the liver and an increasing in the accumulation<br />

of liposomes in tumor vessels. 105 These experiments suggest that optimizing surface<br />

physicochemical properties of nanoparticles to better match the biochemical and physiological<br />

features of tumors may enhance the intratumoral delivery of nanoparticles for systemic<br />

therapeutic approaches.<br />

For conjugation of the negatively charged aptamers to nanoparticles, the surface charge of the<br />

nanoparticle may be important. For example, we believe that direct immobilization of aptamers on<br />

cationic nanoparticles made from PEI may result in formation of aptamer–PEI complex that render<br />

the aptamer ineffective as a targeting molecule (unpublished observation). Therefore, neutral<br />

polymers such as PLA, PLGA or those with a more negative charge, such as polyanhyrides, may<br />

be most suitable for conjugation to aptamers. We have used PLA–PEG block copolymers to<br />

generate aptamer–nanoparticles bioconjugates. 23,108 One approach that may facilitate the use of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 299<br />

a wider array of biomaterials for aptamer targeted drug delivery is through methods of “masking”<br />

the surface charge of the particles. For example, the addition of neutrally charged hydrophilic layer<br />

of PEG on the surface of the nanoparticles may facilitate the use of positively charged materials for<br />

the synthesis of nanoparticles. These cationic nanoparticles are particularly useful for gene delivery<br />

applications and thus may enable efficient targeted gene delivery using aptamers.<br />

16.3.1.4 Surface Modification of Nanoparticles<br />

Nanoparticle surface modification may also be used to engineer their interaction with the<br />

surrounding tissue. These interactions could be positive (i.e., targeting molecules) or negative<br />

(i.e., nonadhesive coatings). The surface modification of nanoparticles is particularly important<br />

because intravenously applied nanoparticles may be captured <strong>by</strong> macrophages before ever reaching<br />

the target site. Therefore, surface modifying particles to render them invisible to macrophages is<br />

essential to making long-circulating nanoparticles. 18,109 The ability to control the biodistribution of<br />

nanoparticles is particularly important for drug carrying nanoparticles because the delivery of drugs<br />

to the normal tissues can lead to toxicity. 16,17<br />

Hydrophilic polymers such as PEG, 18,109 polysaccharides, 110,111 and small molecules 112 can be<br />

conjugated on the surface of nanoparticles to engineer particles with desirable biodistribution<br />

characteristics. For example, to enhance the rate of circulation within the blood and minimize<br />

uptake <strong>by</strong> nondesired cell types, nanoparticles may be coated with polymers such as PEG. 18,109<br />

Various molecular weights and types of PEG (linear or branched) have been used to coat nanoparticles.<br />

113 PEG coatings are also useful for minimizing nanoparticle aggregation and can be used<br />

to prevent clogging of small vasculature and improve size-based targeting. More recently, novel<br />

approaches aimed at conjugating small molecules on nanoparticles using high-throughput methods<br />

have yielded nanoparticle libraries that could be subsequently analyzed for targeted delivery. 112<br />

The use of similar high-throughput approaches has significant potential in optimizing nanoparticle<br />

properties for cancer therapy.<br />

Surface modification of nanoparticles can be achieved in a multistep approach <strong>by</strong> first generating<br />

nanoparticles and subsequently modifying the surface of particles to achieve the desired<br />

characteristics. Alternatively, amphiphilic polymers may be covalently linked prior to generating<br />

nanoparticles to simultaneously control the surface chemistry as well as encapsulate drugs and<br />

COOH<br />

PLA PLA<br />

PEG<br />

Drugs<br />

EDC/NHS NH2 Aptamer<br />

Drugs<br />

FIGURE 16.3 A schematic outlining a conjugation reaction between aptamers and polymer nanoparticles<br />

containing an encapsulated drug. Through incorporating a COOH-terminated, PEG-functionalized surface on<br />

the nanoparticle, NH2-modified aptamers can be easily conjugated using simple aqueous chemistry.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


300<br />

eliminate the need for subsequent chemical modifications once the particle has been synthesized.<br />

This method may provide a more stable coating and better nanoparticle protection in contact with<br />

blood. For example, PLA, poly(caprolactone), and poly(cyanoacrylate) polymers, have been chemically<br />

conjugated to PEG polymers. 18,114,115 We have synthesized nanoparticles from amphiphilic<br />

copolymers composed of lipophilic PLGA and hydrophilic PEG (Figure 16.3) polymers where the<br />

PEG migrates to the surface of the nanoparticles in the presence of an aqueous solution. 18 Asimilar<br />

approach has also been used to generate pegylated PLA nanoparticles using PLA–PEG block<br />

copolymers. 23,108 These particles may be used to extend the nanoparticle residence times in circulation<br />

and enhance accumulation in tumor tissue through “passive targeting” and EPR effect.<br />

In the case of engineering nanoparticles for active targeting, the polymer and its coating should<br />

have functional groups for the attachment of targeting moieties (which may be bound directly to the<br />

nanoparticle surface or though a spacer group). The targeting molecules can enhance the molecular<br />

interaction of the nanoparticles with a subset of cells or tissue.<br />

16.3.2 CONJUGATION STRATEGIES FOR NANOPARTICLE–APTAMER CONJUGATES<br />

Nanotechnology for Cancer Therapy<br />

Covalent conjugation of aptamers to substrates or drug delivery vehicles can be achieved most<br />

commonly through succinimidyl ester–amine chemistry that results in a stable amide linkage 43,43<br />

or through maleimide–thiol chemistry. 23,108,116–120 Potential noncovalent strategies include affinity<br />

interactions (i.e., streptavidin–biotin) and metal coordination (i.e., between a polyhistidine tag at the<br />

end of the aptamer and Ni C2 chelates with immobilized nitrilotriacetic acid on the surface of the<br />

polymer particles). These covalent and noncovalent strategies have been used to immobilize a wide<br />

range of biomolecules, including proteins, enzymes, peptides, and nucleic acids to delivery vehicles.<br />

We believe that covalently linked bioconjugates may result in enhanced stability in physiologic<br />

salt and pH while avoiding the unnecessary addition of biological components (i.e., streptavidin),<br />

thus minimizing immunologic reactions and potential toxicity. For covalent conjugation, the<br />

aptamer is typically modified to carry a terminal primary amine or thiol group that is in turn<br />

conjugated, respectively, to activated carboxylic acid N-hydroxysuccinimide (NHS) ester or maleimide<br />

functional groups present on the surface of drug delivery vehicles. These reactions are carried<br />

out under aqueous conditions with a product yield of 80–90%. 121 One potential difficulty with<br />

maleimide–thiol chemistry is the oxidation of the thiol group attached to aptamers during storage<br />

(formation of S–S bond between two thiol-modified aptamers), resulting in dimers of aptamers that<br />

are not able to participate in the conjugation reaction with the malimide group on particles.<br />

This problem can be partially alleviated <strong>by</strong> using a reducing agent, such as tris(2carboxyethyl)phosphine<br />

(TCEP), b-mercaptoethanol, or dithiothreitol (DTT), during the conjugation<br />

reaction. A potential advantage of using NHS–amine chemistry is that the unreacted<br />

carboxylic acid groups on the particle surface make the particle surface charge (zeta potential)<br />

slightly negative, thus reducing nonspecific interaction between the negatively charged aptamers<br />

and the negative particle surface. Recently, controlled-release nanoparticles generated from<br />

PLA–PEG block copolymer with a terminal carboxylic acid group attached to the PEG were<br />

conjugated with primary-amine-terminated aptamers. 23,103 In this case, the hydrophilic PEG<br />

group facilitated the presentation of the carboxylic acid on the particle surface for conversion to<br />

activated carboxylic acid NHS ester and conjugation to the primary-amine-modified aptamers.<br />

The conjugation of aptamers to nanoparticles can be qualitatively confirmed <strong>by</strong> fluorescent<br />

microscopy or flow cytometry through the use of fluorescent probes, such as fluorescein isothiocyanate<br />

(FITC), that are conjugated directly to the aptamers or indirectly to complementary<br />

oligonucleotides that hybridize to the aptamers. 23 Alternatively, analytical approaches, such as<br />

x-ray photoemission spectroscopy (XPS), may be used for characterization of the nanoparticle<br />

surface to confirm the extent of conjugation. The presence of a hydrocarbon spacer group<br />

between the nanoparticle surface and the aptamer should improve the probability of interaction<br />

between the aptamer and its target. Furthermore, a consistent density of the aptamer on the surface<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 301<br />

of nanoparticles can potentially be achieved <strong>by</strong> utilizing an excess molar amount of aptamer relative to<br />

the reactive group on the nanoparticle surface during the conjugation reactions. However, the optimal<br />

density of targeting molecule on nanoparticle surface may need to be experimentally determined. 122<br />

We have used the covalent conjugation approach to demonstrate a proof-of-concept for nanoparticle–aptamer<br />

bioconjugates that target the PSMA on the surface of prostate cancer cells and are<br />

taken up <strong>by</strong> cells that express the PSMA protein specifically and efficiently. 23 We have also shown<br />

using a microfluidic system that these nanoparticle–aptamer conjugates are capable of binding to<br />

their target cells under flow conditions suggesting their suitability for in vivo targeted drug-delivery<br />

applications. 108 Most recently, we have demonstrated the in vivo efficacy of docetaxel-encapsulated<br />

nanoparticle–aptamer conjugates using a xenograft prostate cancer nude mouse model. 14<br />

These approaches have paved the way for the use of aptamers for targeted delivery of drug<br />

encapsulated nanoparticles to a myriad of human cancers.<br />

16.4 APTAMER–DRUG CONJUGATES FOR TARGETED CANCER THERAPY<br />

The clinical utility of radioimmunoconjugates and chemoimmunoconjugates in cancer treatment is<br />

well documented with several examples of each in clinical practice or under development at this<br />

time. 123–125 More recently aptamer conjugates for cancer therapeutic and diagnostic applications,<br />

14,23,108,116,118–120,126 biosensors, 127–129 flow cytometry, 130 ELISAs, 131,132 and capillary<br />

electrophoresis for separation technology have been discribed. 40,133 As noted in Section 16.2,<br />

aptamers have distinct advantages over antibodies, and, for the purpose of targeted delivery of<br />

toxins, their greatest advantage lies in their relatively smaller size and remarkable stability.<br />

The smaller size of aptamers, as compared to antibodies, allows for a more efficient tumor<br />

penetration of aptamers, translating to a more efficient delivery of a cytotoxic payload to tumor<br />

cells. 119 Our research group is interested in the use of aptamer–toxin conjugates for therapeutic<br />

applications, 126 and we believe that aptamer–toxin conjugates may result in future novel<br />

therapeutic applications in oncology and other areas of medicine.<br />

For the preparation of the aptamer–drug conjugates, conventional conjugation strategies that<br />

have been used for the preparation of immunoconjugates may also be employed with some modifications.<br />

125 For targeted radiotherapy, beta-emitting radioisotopes, such as 131 I, 90 Y, or 67 Cu, or<br />

alpha-emitting radioisotopes such as 213 Bi or 211 At, may be used to develop radioaptamerconjugates.<br />

These radioisotopes may be linked directly to aptamers or immobilized through metal<br />

chelators attached to aptamers. For targeted chemotherapy, drug molecules (i.e., doxorubicin,<br />

calicheamicin, or auristatin) may be covalently coupled to one or more sites on the aptamer<br />

using simple chemistry (i.e., formation of amide, disulfide, or hydrazone bond). 134,135 The development<br />

of aptamer conjugates is facilitated <strong>by</strong> site-selective functionalization of aptamers at their 3 0 -<br />

and/or 5 0 -terminus, making it possible to reproducibly attach an exact number of drugs in a sitespecific<br />

manner. 23,108,136 Aptamers may also be functionalized within the oligonucleotide chain<br />

using modified nucleotides with desired functional groups; however, this latter approach may<br />

negatively impact the aptamer conformation and function. In the case of conjugation of toxins to<br />

antibodies, the reaction is most commonly not site-directed, and this nonspecific conjugation<br />

approach may negatively impact the binding characteristics of antibodies. 136 Furthermore, unlike<br />

antibodies, the desired linkers or functional groups on aptamers for conjugation on other applications<br />

can be easily introduced during the chemical synthesis of these molecules. We have<br />

schematically demonstrated several proposed conjugation strategies between aptamers and drugs<br />

(Figure 16.4). Typical functional groups modifiable at the 5 0 -or3 0 -ends of aptamers are mostly<br />

nucleophiles, including amine (–NH2), sulfhydryl (–SH), and hydrazide (–C(O)NHNH2). Appropriate<br />

functional groups on the drugs for immobilization on aptamers may be carboxylic acid (for amide bond<br />

formation with amine-modified aptamers), maleimide or unsaturated carbonyl (for thioether formation<br />

with sulfhydryl groups on aptamers), activated disulfide (for example, S-pyridyl disulfide) for disulfide<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


302<br />

3'<br />

5'<br />

X<br />

3' 3'<br />

X Y<br />

5' 5'<br />

X X<br />

aptamer<br />

Site-specific conjugation<br />

3' 3' 3' PEG 3'<br />

5' 5' 5' 5'<br />

(a) (b) (c) (d)<br />

formation with sulfhydryl groups on aptamers, and ketone for hydrazone formation with hydrazide on<br />

aptamers. 136<br />

The covalent bonds formed through the aforementioned strategies are all cleavable in vivo,<br />

allowing for the active drug to be released for therapeutic efficacy. The choice of conjugation<br />

strategy may allow for a spatial control of this release and improved clinical efficacy. In addition, <strong>by</strong><br />

altering the size of aptamer–drug complex, the pharmacokinetics properties of the conjugate may<br />

be controlled, allowing for temporal control of the drug circulation and clearance. For example, a longer<br />

circulation time and reduced renal clearance may be achieved <strong>by</strong> linking the aptamer–drug conjugates<br />

to PEG polymers to increase the size of the complex above the threshold for renal clearance. 137<br />

This conjugate form may result in improved drug delivery to cancers as a result of increased systemic<br />

circulation time. Intracellular versus extracellular drug release also have unique sets of design<br />

considerations. 138 In the case of the disulfide linkage approach, the aptamer–drug conjugate is expected<br />

to be stable in plasma during circulation with subsequent intracellular cleavage of disulfide bond and<br />

drug release due to high intracellular concentration of glutathione (GSH). 139 Hydrazone linkage also<br />

has favorable properties for conjugation of drugs to aptamers, allowing for a pH sensitive release of the<br />

drug in a relatively more acidic environment of the tumor tissues. 134,140,141 Therefore, disulfide and<br />

hydrazone linkages may be considered when developing aptamer conjugates for cancer therapeutics.<br />

We have recently developed a novel strategy for generating aptamer–drug conjugates <strong>by</strong><br />

intercalating drugs between nucleic acid bases of the aptamer (Figure 16.5). 126 Using doxorubicin<br />

(dox) as a model drug and the A10 PSMA aptamer as model aptamer, we have developed a stable<br />

aptamer–dox conjugate through physical interactions. The resulting physical conjugate demonstrated<br />

the following characteristics: (1) high discrimination between cancer cells expressing<br />

target antigen and cells lacking target antigen; and (2) efficient delivery of anti-cancer drug to<br />

target cancer cells. 126 Specifically, we developed aptamer–dox conjugates and demonstrated the<br />

stability of this system in vitro. Using LNCaP prostate cancer cell lines that express the PSMA<br />

antigen and PC3 prostate cancer cell lines that do not express the PSMA protein, we demonstrated a<br />

differential uptake of the aptamer–dox conjugate <strong>by</strong> LNCaP cells, translating to a more efficient<br />

cellular toxicity and anti-cancer efficacy in vitro (Figure 16.6). These results suggest that physical<br />

conjugate strategies such as intercalation may serve as a novel approach for developing<br />

aptamer–drug therapeutics.<br />

linker<br />

X,Y= functional groups<br />

−NH 2 , −SH<br />

−C(O)NHNH 2, etc.<br />

Nanotechnology for Cancer Therapy<br />

Z Z<br />

Drugs<br />

Z = −COOH, Maleimide<br />

−C(O)R<br />

FIGURE 16.4 Available conjugation strategies between end-functionalized aptamers and drug molecules.<br />

Panels (a) through (d) demonstrate a series of single- and double-conjugation strategies.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 303<br />

PSMA aptamer<br />

+<br />

Doxorubicin<br />

intercalation<br />

Physical conjugate<br />

FIGURE 16.5 A schematic diagram outlining the formation of physical conjugate between aptamers and<br />

drugs capable of intercalating into base pairs of aptamers. Doxorubicin, an anti-cancer drug, can be intercalated<br />

into a PSMA aptamer to form a 1:1 complex.<br />

FIGURE 16.6 The confocal laser scanning microscopy images of LNCaP (a) and PC3 cells (b) after<br />

treatments of the aptamer–doxorubicin physical conjugate (1.5 mM) for 2 h.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


304<br />

In addition to small molecule and radioisotopes, bio therapeutics such as small interfering<br />

RNAs (siRNA) 142–144 and protein-based toxins may also be conjugated to aptamers. For<br />

example, siRNAs can be coupled to aptamers through disulfide bond, allowing the release of<br />

siRNAs in the active form after cytosolic uptake <strong>by</strong> the targeted cells. A conjugate of aptamer–<br />

siRNA may represent a novel class of therapeutics with widespread application in medicine. 116<br />

16.5 APTAMER–NANOPARTICLE CONJUGATES FOR PROTEIN DETECTION<br />

Quantum dots (QDs) are nanometer-sized semiconductor crystals with tunable fluorescence that<br />

can be developed as sensitive biosensors or probes for in vivo imaging and diagnostic applications.<br />

145–147 Unlike conventional small molecule fluorophores and biological fluorophores,<br />

such as the green fluorescence protein (GFP), QDs possess several favorable characteristics,<br />

including: (1) greatly improved photostability (lack of photobleaching) that enable real time<br />

imaging of biomarkers over an extended period of time; (2) a broad excitation wavelength and a<br />

narrow emission wavelength enabling excitation of multiple QDs simultaneously using a single<br />

laser wavelength; (3) bright fluorescence with high quantum yield; and (4) ability to functionalize<br />

the surface of QDs with various molecules ranging from small molecules to antibodies and aptamers<br />

for desired targeted applications. Taken together, these favorable characteristics have made<br />

QDs an emerging class of imaging probes with broad applications in medicine.<br />

Metallic nanoparticles, such as gold-nanoparticles (Gd-NPs), are capable of absorbing or scattering<br />

light and are now used in many biological applications, particularly visualization of cellular<br />

and tissue components <strong>by</strong> electron microscopy. 148,149 More recently, it has been reported that<br />

antibody–Gd-NP conjugates 150,151 can detect proteins 152 or viruses 153 in solution through surface<br />

plasmon resonance effect, which results in color change depending on the status of aggregation of<br />

the particles. This simple phenomenon together with the fact that Gd-NPs are biocompatible makes<br />

Gd-NPs useful detection probes of various in vivo applications. 154<br />

The surface functionalization of QDs and Gd-NPs with specific targeting molecules may result<br />

in the development of targeted imaging modalities with widespread applications in research<br />

and medicine.<br />

16.5.1 APTAMER–QD CONJUGATES FOR THE DETECTION OF PROTEINS<br />

A QD–aptamer beacon for the detection of thrombin was recently reported, 155 and the principle of<br />

this detection system is schematically illustrated (Figure 16.7). Briefly, QDs were surface<br />

Q<br />

Q<br />

QD<br />

Q<br />

Q<br />

thrombin<br />

Quenched fluorescence Enhanced fluorescence<br />

QD<br />

Nanotechnology for Cancer Therapy<br />

+<br />

Q<br />

Quencher−oligoDNA<br />

FIGURE 16.7 A schematic diagram outlining QD–aptamer beacon system for the detection of thrombin.<br />

Target protein replaces the bound quencher–oligoDNA <strong>by</strong> forming stable complex with aptamers on a QD,<br />

resulting in recovery of a native QD’s fluorescence.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 305<br />

functionalized with an aptamer known to bind to the thrombin protein to develop a thrombindetecting<br />

QD–aptamer beacon. A second oligonucleotide, having complementary sequences to the<br />

thrombin aptamer and bearing a quencher, was synthesized and utilized to quench the system in the<br />

absence of thrombin. When the quencher oligonucleotide was incubated with the QD–aptamer<br />

conjugate in the absence of thrombin, the hybridization of the quencher oligonucleotide and the<br />

thrombin aptamer proximated the quencher molecules to the surface of the QD, resulting in<br />

quenching effect through fluorescence resonance energy transfer (FRET). 156,157 In the presence<br />

of thrombin (1 mM), the quencher oligonucleotide is displaced, resulting in restoration of the<br />

fluorescence of the QD–aptamer conjugate. Using this QD–aptamer beacon system, a 1 mM concentration<br />

of thrombin could be detected specifically. We postulate that multiplexed detection of a<br />

panel of protein targets may be performed in parallel <strong>by</strong> formation of QD–aptamer conjugates from<br />

different aptamers, and QDs possessing different emitting fluorescence profile. 158,159 In addition,<br />

considering a growing number of tumor-specific or tumor-associated aptamers that are available<br />

today, it may be increasingly possible to develop more specific and more efficient cancer diagnostic<br />

probes for a myriad of important cancers. We have recently developed a QD–aptamer conjugate<br />

using the A10 PSMA aptamer and have demonstrated the differential binding and uptake of these<br />

vehicles <strong>by</strong> cells that express the PSMA antigen (unpublished results).<br />

16.5.2 APTAMER–GOLD-NANOPARTICLE CONJUGATES FOR THE DETECTION OF PROTEINS<br />

A highly specific sensing system for the detection of platelet-derived growth factor (PDGF) has<br />

been developed using PDGF-specific, aptamer–conjugated gold-nanoparticles. 160 The principle of<br />

the detection using the conjugate is illustrated in Figure 16.8. The binding of soluble PDGF protein<br />

to Gd-NP–aptamer conjugates results in a change in the color of the solution due to the formation of<br />

aggregates between the Gd-NP conjugates. In the absence of PDGF, the conjugates are dispersed<br />

separately in solution (red), but in the presence of PDGF the Gd-NP aggregate closely to an interparticle<br />

distances less than the average diameter of each nanoparticle, causing a dipole–dipole<br />

coupling between particles as well as increased scattering, resulting in a measurable color<br />

change (purple). 161–163 Using this detection system, Gd-NP–aptamer conjugate could detect nanomolar<br />

concentrations of PDGF in solution. Because Gd-NPs are biocompatible, easy to develop,<br />

and easy to functionalize with aptamers, it is expected that bioconjugates utilizing aptamers and<br />

Gd-NPs as target molecular probes may result in novel cancer diagnostic modalities in the future.<br />

Gd-NP−Aptamer conjugates in dispersion<br />

target protein<br />

Gd-NP−Aptamer conjugates after aggregation<br />

FIGURE 16.8 A schematic representation of the aggregation of gold-nanoparticle–aptamer conjugates in the<br />

presence of target proteins. Aggregation causes color change from red to purple.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


306<br />

16.6 CONCLUSION<br />

Bioconjugates comprising nanoparticles and aptamers represent a potentially powerful tool for<br />

developing novel diagnostic and therapeutic modalities for cancer detection and treatment. As<br />

drug delivery vehicles for cancer therapy, nanoparticle–aptamer bioconjugates can be designed<br />

to target and be taken up <strong>by</strong> cancer cells for targeted delivery and controlled release of chemotherapeutic<br />

drugs over an extended time directly at the site of tumors. The successful achievement<br />

of this goal requires the isolation of aptamers that bind to the extracellular domain of antigens<br />

expressed exclusively or preferentially on the plasma membrane of cancer cells or on the extra<br />

cellular matrices of tumor tissue. In addition, nanoparticles would have to be designed with the<br />

optimized properties that facilitate targeting and delivery of the drugs to the desired tissues while<br />

avoiding uptake <strong>by</strong> the mononuclear phagocytic system in the body.<br />

The targeted delivery of chemotherapeutic drugs for cancer therapy may minimize their side<br />

effects and enhance their cytotoxicity to cancer cells, resulting in a better clinical outcome. We<br />

anticipate that the combination of controlled release technology and targeted approaches may<br />

represent a viable approach for achieving this goal. One major clinical advantage of targeted<br />

drug-encapsulated nanoparticle conjugates over drugs that are directly linked to a targeting<br />

moiety is that large amounts of chemotherapeutic drug may be delivered to cancer cells per each<br />

delivery and biorecognition event. Another advantage would be the ability to simultaneously<br />

deliver two or more chemotherapeutic drugs and release each in a predetermined manner, thus<br />

resulting in effective combination chemotherapy, which is common for the management of many<br />

cancers. Antibodies and peptides have been widely used for the targeted delivery of drug encapsulated<br />

nanoparticles; however, the translation of these vehicles into clinical practice has lagged<br />

behind our advances in the laboratory. This has been largely due to the immunogenicity and<br />

nonspecific uptake of nanoparticle–antibody bioconjugates <strong>by</strong> nontargeted cells, resulting in<br />

toxicity or poor efficacy of these conjugates. Nanoparticle–aptamer bioconjugates face some of<br />

the same challenges of nonspecific uptake after systemic administration and thus must be engineered<br />

with surface physicochemical characteristics to avoid toxicity to nontargeted cells. We<br />

believe that optimal particle size and surface properties to sufficiently decrease the rate of nonspecific<br />

particle uptake while achieving successful targeting must be determined experimentally on a<br />

case-<strong>by</strong>-case basis, as this also depends on the polymer system, the drug being encapsulated and the<br />

tumor microenvironment including its vascularity. However, the advantage of nanoparticle–<br />

aptamer bioconjugates over their antibody conjugate counterparts lies in the ease of aptamer<br />

isolation, lack of immunogenicity, and a decreased batch-to-batch variability attributable to the<br />

chemical nature of aptamer synthesis and production, which can facilitate the translation of optimally<br />

engineered nanoparticle–aptamer bioconjugates into clinical practice.<br />

ACKNOWLEDGMENTS<br />

The authors thank Jack Szostak, Benjamin Teply, Jeffrey Karp, Jianjun Cheng, Chris Cannizarro,<br />

Etgar Levy-Nissenbaum, and Andrej Luptak for helpful discussions and review of this manuscript.<br />

This work was supported <strong>by</strong> NIH/NCI CA119349, NIH/NIBIB EB003647, and <strong>by</strong> the David Koch<br />

Cancer Research Fund.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Peppas, N. A., Intelligent therapeutics: biomimetic systems and nanotechnology in drug delivery,<br />

Adv. Drug Deliv. Rev., 56(11), 1529–1531, 2004.<br />

2. Langer, R. and Tirrell, D. A., Designing materials for biology and medicine, Nature, 428(6982),<br />

487–492, 2004.<br />

3. Langer, R., Drug delivery and targeting, Nature, 392(6679), 5–10, 1998.<br />

4. Ferrari, M., Cancer nanotechnology: opportunities and challenges, Nat. Rev. Cancer, 5(3), 161–171, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 307<br />

5. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: mechanism of tumoritropic accumulation of proteins and the antitumor agent<br />

smancs, Cancer Res., 46(12 Pt 1), 6387–6392, 1986.<br />

6. Maeda, H. and Matsumura, Y., Tumoritropic and lymphotropic principles of macromolecular drugs,<br />

Crit. Rev. Ther. Drug Carrier Syst., 6(3), 193–210, 1989.<br />

7. Maeda, H., The enhanced permeability and retention (EPR) effect in tumor vasculature: the key role<br />

of tumor-selective macromolecular drug targeting, Adv. Enzyme Regul., 41, 189–207, 2001.<br />

8. Yuan, F., Dellian, M., Fukumura, D., Leunig, M., Berk, D. A., Torchilin, V. P., and Jain, R. K.,<br />

Vascular permeability in a human tumor xenograft: molecular size dependence and cutoff size,<br />

Cancer Res., 55(17), 3752–3756, 1995.<br />

9. Kong, G., Braun, R. D., and Dewhirst, M. W., Hyperthermia enables tumor-specific nanoparticle<br />

delivery: effect of particle size, Cancer Res., 60(16), 4440–4445, 2000.<br />

10. Yuan, F., Leunig, M., Huang, S. K., Berk, D. A., Papahadjopoulos, D., and Jain, R. K., Microvascular<br />

permeability and interstitial penetration of sterically stabilized (stealth) liposomes in a human tumor<br />

xenograft, Cancer Res., 54(13), 3352–3356, 1994.<br />

11. Brigger, I., Morizet, J., Laudani, L., Aubert, G., Appel, M., Velasco, V., Terrier-Lacombe, M. J.<br />

et al., Negative preclinical results with stealth nanospheres-encapsulated Doxorubicin in an orthotopic<br />

murine brain tumor model, J. Control Release, 100(1), 29–40, 2004.<br />

12. Brigger, I., Dubernet, C., and Couvreur, P., Nanoparticles in cancer therapy and diagnosis, Adv. Drug<br />

Deliv. Rev., 54(5), 631–651, 2002.<br />

13. Anderson, D. G., Peng, W., Akinc, A., Hossain, N., Kohn, A., Padera, R., Langer, R., and Sawicki,<br />

J. A., A polymer library approach to suicide gene therapy for cancer, Proc. Natl Acad. Sci. U.S.A.,<br />

101(45), 16028–16033, 2004.<br />

14. Farokhzad, O. C., Cheng, J., Teply, B. A., Sherifi, I., Jon, S., Kantoff, P. W., Richie, J. P., and Langer,<br />

R., Targeted nanoparticle-aptamer bioconjugates for cancer chemotherapy in vivo, Proc. Natl Acad.<br />

Sci. U.S.A., 103(16), 6315–6320, <strong>2006</strong>.<br />

15. Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancer therapy,<br />

Adv. Drug Del. Rev., 56(11), 1649–1659, 2004.<br />

16. Demoy, M., Gibaud, S., Andreux, J. P., Weingarten, C., Gouritin, B., and Couvreur, P., Splenic<br />

trapping of nanoparticles: complementary approaches for in situ studies, Pharm. Res., 14(4),<br />

463–468, 1997.<br />

17. Gibaud, S., Andreux, J. P., Weingarten, C., Renard, M., and Couvreur, P., Increased bone marrow<br />

toxicity of doxorubicin bound to nanoparticles, Eur. J. Cancer, 30A(6), 820–826, 1994.<br />

18. Gref, R., Minamitake, Y., Peracchia, M. T., Trubetskoy, V., Torchilin, V., and Langer, R., Biodegradable<br />

long-circulating polymeric nanospheres, Science, 263(5153), 1600–1603, 1994.<br />

19. Monsky, W. L., Fukumura, D., Gohongi, T., Ancukiewcz, M., Weich, H. A., Torchilin, V. P., Yuan,<br />

F., and Jain, R. K., Augmentation of transvascular transport of macromolecules and nanoparticles in<br />

tumors using vascular endothelial growth factor, Cancer Res., 59(16), 4129–4135, 1999.<br />

20. Guillemard, V. and Saragovi, H. U., Novel approaches for targeted cancer therapy, Curr. Cancer<br />

Drug Targets, 4(4), 313–326, 2004.<br />

21. Zhang, X. Q., Wang, X. L., Zhang, P. C., Liu, Z. L., Zhuo, R. X., Mao, H. Q., and Leong, K. W.,<br />

Galactosylated ternary DNA/polyphosphoramidate nanoparticles mediate high gene transfection<br />

efficiency in hepatocytes, J. Control Release, 102(3), 749–763, 2005.<br />

22. Hicke, B. J. and Stephens, A. W., Escort aptamers: a delivery service for diagnosis and therapy,<br />

J. Clin. Invest., 106(8), 923–938, 2000.<br />

23. Farokhzad, O. C., Jon, S., Khademhosseini, A., Tran, T. N., Lavan, D. A., and Langer, R., Nanoparticle-aptamer<br />

bioconjugates: a new approach for targeting prostate cancer cells, Cancer Res.,<br />

64(21), 7668–7672, 2004.<br />

24. Sullenger, B. A., Gallardo, H. F., Ungers, G. E., and Gilboa, E., Overexpression of TAR sequences<br />

renders cells resistant to human immunodeficiency virus replication, Cell, 63(3), 601–608, 1990.<br />

25. Ellington, A. D. and Szostak, J. W., In vitro selection of RNA molecules that bind specific ligands,<br />

Nature, 346(6287), 818–822, 1990.<br />

26. Tuerk, C. and Gold, L., Systematic evolution of ligands <strong>by</strong> exponential enrichment: RNA ligands to<br />

bacteriophage T4 DNA polymerase, Science, 249(4968), 505–510, 1990.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


308<br />

Nanotechnology for Cancer Therapy<br />

27. James, J. S. and Dubs, G., FDA approves new kind of lymphoma treatment. Food and drug administration,<br />

AIDS Treat. News,No 284 2–3, 1997.<br />

28. Foss, F. M., DAB(389)IL-2 (ONTAK): a novel fusion toxin therapy for lymphoma, Clin. Lymphoma,<br />

1(2), 110–116, discussion 117, 2000.<br />

29. Harris, L. J., Larson, S. B., Hasel, K. W., Day, J., Greenwood, A., and McPherson, A., The threedimensional<br />

structure of an intact monoclonal antibody for canine lymphoma, Nature, 360(6402),<br />

369–372, 1992.<br />

30. Harris, L. J., Skaletsky, E., and McPherson, A., Crystallographic structure of an intact IgG1 monoclonal<br />

antibody, J. Mol. Biol., 275(5), 861–872, 1998.<br />

31. Marks, J. D., Selection of internalizing antibodies for drug delivery, Methods Mol. Biol., 248,<br />

201–208, 2004.<br />

32. Marks, J. D., Ouwehand, W. H., Bye, J. M., Finnern, R., Gorick, B. D., Voak, D., Thorpe, S. J.,<br />

Hughes-Jones, N. C., and Winter, G., Human antibody fragments specific for human blood group<br />

antigens from a phage display library, Biotechnology (N Y), 11(10), 1145–1149, 1993.<br />

33. Doggrell, S. A., Pegaptanib: the first antiangiogenic agent approved for neovascular macular<br />

degeneration, Expert Opin. Pharmacother., 6(8), 1421–1423, 2005.<br />

34. Fitzwater, T. and Polisky, B., A SELEX primer, Methods Enzymol., 267, 275–301, 1996.<br />

35. White, R. R., Sullenger, B. A., and Rusconi, C. P., Developing aptamers into therapeutics, J. Clin.<br />

Invest., 106(8), 929–934, 2000.<br />

36. Huang, D. B., Vu, D., Cassiday, L. A., Zimmerman, J. M., Maher, L. J., and Ghosh, G., Crystal<br />

structure of NF-kappaB (p50)2 complexed to a high-affinity RNA aptamer, Proc. Natl Acad. Sci.<br />

U.S.A., 100(16), 9268–9273, 2003.<br />

37. Hermann, T. and Patel, D. J., Adaptive recognition <strong>by</strong> nucleic acid aptamers, Science, 287(5454),<br />

820–825, 2000.<br />

38. Drolet, D. W., Nelson, J., Tucker, C. E., Zack, P. M., Nixon, K., Bolin, R., Judkins, M. B. et al.,<br />

Pharmacokinetics and safety of an anti-vascular endothelial growth factor aptamer (NX1838) following<br />

injection into the vitreous humor of rhesus monkeys, Pharm. Res., 17(12), 1503–1510, 2000.<br />

39. Jenison, R. D., Gill, S. C., Pardi, A., and Polisky, B., High-resolution molecular discrimination <strong>by</strong><br />

RNA, Science, 263(5152), 1425–1429, 1994.<br />

40. Tombelli, S., Minunni, M., and Mascini, M., Analytical applications of aptamers, Biosens. Bioelectron.,<br />

20(12), 2424–2434, 2005.<br />

41. Charlton, J., Sennello, J., and Smith, D., In vivo imaging of inflammation using an aptamer inhibitor<br />

of human neutrophil elastase, Chem. Biol., 4(11), 809–816, 1997.<br />

42. Beigelman, L., McSwiggen, J. A., Draper, K. G., Gonzalez, C., Jensen, K., Karpeisky, A. M., Modak,<br />

A. S. et al., Chemical modification of hammerhead ribozymes. Catalytic activity and nuclease<br />

resistance, J. Biol. Chem., 270(43), 25702–25708, 1995.<br />

43. Aurup, H., Williams, D. M., and Eckstein, F., 2&prime;-Fluoro- and 2&prime;-amino-2&prime;deoxynucleoside<br />

5&prime;-triphosphates as substrates for T7 RNA polymerase, Biochemistry,<br />

31(40), 9636–9641, 1992.<br />

44. Pieken, W. A., Olsen, D. B., Benseler, F., Aurup, H., and Eckstein, F., Kinetic characterization of ribonuclease-resistant<br />

2&prime;-modified hammerhead ribozymes, Science, 253(5017), 314–317, 1991.<br />

45. Eulberg, D. and Klussmann, S., Spiegelmers: biostable aptamers, Chem. biochem., 4(10), 979–983,<br />

2003.<br />

46. Schmidt, K. S., Borkowski, S., Kurreck, J., Stephens, A. W., Bald, R., Hecht, M., Friebe, M.,<br />

Dinkelborg, L., and Erdmann, V. A., Application of locked nucleic acids to improve aptamer in<br />

vivo stability and targeting function, Nucleic Acids Res., 32(19), 5757–5765, 2004.<br />

47. Tucker, C. E., Chen, L. S., Judkins, M. B., Farmer, J. A., Gill, S. C., and Drolet, D. W., Detection and<br />

plasma pharmacokinetics of an anti-vascular endothelial growth factor oligonucleotide-aptamer<br />

(NX1838) in rhesus monkeys, J Chromatogr. B Biomed. Sci. Appl., 732(1), 203–212, 1999.<br />

48. Gold, L. and Tuerk, L., Methods of Producing Nucleic Acid Ligands, U.S.A., 1997.<br />

49. Chelliserrykattil, J. and Ellington, A. D., Evolution of a T7 RNA polymerase variant that transcribes<br />

2&prime;-O-methyl RNA, Nat. Biotechnol., 22(9), 1155–1160, 2004.<br />

50. Burmeister, P. E., Lewis, S. D., Silva, R. F., Preiss, J. R., Horwitz, L. R., Pendergrast, P. S.,<br />

McCauley, T. G. et al., Direct in vitro selection of a 2&prime;-O-methyl aptamer to VEGF,<br />

Chem. Biol., 12(1), 25–33, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 309<br />

51. Lee, J. F., Hesselberth, J. R., Meyers, L. A., and Ellington, A. D., Aptamer database, Nucleic Acids<br />

Res., 32, D95–D100, 2004.<br />

52. Gragoudas, E. S., Adamis, A. P., Cunningham, E. T., Feinsod, M., and Guyer, D. R., Pegaptanib for<br />

neovascular age-related macular degeneration, N Engl. J. Med., 351(27), 2805–2816, 2004.<br />

53. Vinores, S. A., Technology evaluation: pegaptanib. Eyetech/Pfizer, Curr. Opin. Mol. Ther., 5(6),<br />

673–679, 2003.<br />

54. Pestourie, C., Tavitian, B., and Duconge, F., Aptamers against extracellular targets for in vivo<br />

applications, Biochimie, 87(9-10), 921–930, 2005.<br />

55. McNeel, D. G., Eickhoff, J., Lee, F. T., King, D. M., Alberti, D., Thomas, J. P., Friedl, A. et al., Phase<br />

I trial of a monoclonal antibody specific for alphavbeta3 integrin (MEDI-522) in patients with<br />

advanced malignancies, including an assessment of effect on tumor perfusion, Clin. Cancer Res.,<br />

11(21), 7851–7860, 2005.<br />

56. Mi, J., Zhang, X., Giangrande, P. H., McNamara, J. O., Nimjee, S. M., Sarraf-Yazdi, S., Sullenger,<br />

B. A., and Clary, B. M., Targeted inhibition of alphavbeta3 integrin with an RNA aptamer impairs<br />

endothelial cell growth and survival, Biochem. Biophys. Res. Commun., 338(2), 956–963, 2005.<br />

57. Chen, C. H., Chernis, G. A., Hoang, V. Q., and Landgraf, R., Inhibition of heregulin signaling <strong>by</strong> an<br />

aptamer that preferentially binds to the oligomeric form of human epidermal growth factor receptor-<br />

3, Proc. Natl Acad. Sci. U.S.A., 100(16), 9226–9231, 2003.<br />

58. Lupold, S. E., Hicke, B. J., Lin, Y., and Coffey, D. S., Identification and characterization of nucleasestabilized<br />

RNA molecules that bind human prostate cancer cells via the prostate-specific membrane<br />

antigen, Cancer Res., 62(14), 4029–4033, 2002.<br />

59. Deng, J. S., Ballou, B., and Hofmeister, J. K., Internalization of anti-nucleolin antibody into viable<br />

HEp-2 cells, Mol. Biol. Rep., 23(3-4), 191–195, 1996.<br />

60. Christian, S., Pilch, J., Akerman, M. E., Porkka, K., Laakkonen, P., and Ruoslahti, E., Nucleolin<br />

expressed at the cell surface is a marker of endothelial cells in angiogenic blood vessels, J. Cell Biol.,<br />

163(4), 871–878, 2003.<br />

61. Dapic, V., Abdomerovic, V., Marrington, R., Peberdy, J., Rodger, A., Trent, J. O., and Bates, P. J.,<br />

Biophysical and biological properties of quadruplex oligodeoxyribonucleotides, Nucleic Acids Res.,<br />

31(8), 2097–2107, 2003.<br />

62. Barnhart, K. M., Laber, D. A., Bates, P. J., Trent, J. O., and Miller, D. M., In AGRO100: The translation<br />

from lab to clinic of a tumor-targeted nucleic acid aptamer. Proceedings from the American Society<br />

for Clinical Oncology Meeting, American Society for Clinical Oncology, New Orleans, LA, 2004.<br />

63. Laber, D. A., Sharma, V. R., Bhupalam, L., Taft, B., Hendler, F. J., and Barnhart, K. M. In Update on<br />

the First Phase I Study of AGRO100 in Advanced Cancer, American Society of Clinical Oncology<br />

Annual Meeting, Abstract 3064, 2005.<br />

64. Girvan, A. C., Barve, S. S., Thongboonkerd, V., Klein, J. B., and Pierce, W. M., In Inhibition of<br />

NFkB signaling <strong>by</strong> an oligonucleotide aptamer, Proceedings of the American Association of Cancer<br />

Research, 2004.<br />

65. Jeong, S., Eom, T., Kim, S., Lee, S., and Yu, J., In vitro selection of the RNA aptamer against the Sialyl<br />

Lewis X and its inhibition of the cell adhesion,Biochem. Biophys. Res. Commun., 281(1),237–243,2001.<br />

66. Santulli-Marotto, S., Nair, S. K., Rusconi, C., Sullenger, B., and Gilboa, E., Multivalent RNA<br />

aptamers that inhibit CTLA-4 and enhance tumor immunity, Cancer Res., 63(21), 7483–7489, 2003.<br />

67. Hicke, B. J., Marion, C., Chang, Y. F., Gould, T., Lynott, C. K., Parma, D., Schmidt, P. G., and<br />

Warren, S., Tenascin-C aptamers are generated using tumor cells and purified protein, J. Biol. Chem.,<br />

276(52), 48644–48654, 2001.<br />

68. Blank, M., Weinschenk, T., Priemer, M., and Schluesener, H., Systematic evolution of a DNA<br />

aptamer binding to rat brain tumor microvessels. selective targeting of endothelial regulatory<br />

protein pigpen, J. Biol. Chem., 276(19), 16464–16468, 2001.<br />

69. Santini, J. T., Cima, M. J., and Langer, R., A controlled-release microchip, Nature, 397(6717),<br />

335–338, 1999.<br />

70. Edwards, D. A., Hanes, J., Caponetti, G., Hrkach, J., Ben-Jebria, A., Eskew, M. L., Mintzes, J.,<br />

Deaver, D., Lotan, N., and Langer, R., Large porous particles for pulmonary drug delivery, Science,<br />

276(5320), 1868–1871, 1997.<br />

71. LaVan, D. A., McGuire, T., and Langer, R., Small-scale systems for in vivo drug delivery, Nature<br />

Biotechnol., 21(10), 1184–1191, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


310<br />

Nanotechnology for Cancer Therapy<br />

72. Elisseeff, J., McIntosh, W., Fu, K., Blunk, B. T., and Langer, R., Controlled-release of IGF-I and<br />

TGF-beta1 in a photopolymerizing hydrogel for cartilage tissue engineering, J. Orthop. Res., 19(6),<br />

1098–1104, 2001.<br />

73. Tanabe, K., Regar, E., Lee, C. H., Hoye, A., van der Giessen, W. J., and Serruys, P. W., Local drug<br />

delivery using coated stents: new developments and future perspectives, Curr. Pharm. Des., 10(4),<br />

357–367, 2004.<br />

74. Ebrahim, S., Peyman, G. A., and Lee, P. J., Applications of liposomes in ophthalmology, Surv.<br />

Ophthalmol., 50(2), 167–182, 2005.<br />

75. Haak, T., New developments in the treatment of type 1 diabetes mellitus, Exp. Clin. Endocrinol.<br />

Diabetes, 107(suppl. 3), S108–S113, 1999.<br />

76. Guerin, C., Olivi, A., Weingart, J. D., Lawson, H. C., and Brem, H., Recent advances in brain tumor<br />

therapy: local intracerebral drug delivery <strong>by</strong> polymers, Invest. New Drugs, 22(1), 27–37, 2004.<br />

77. Jiang, W., Gupta, R. K., Deshpande, M. C., and Schwendeman, S. P., Biodegradable poly(lactic-coglycolic<br />

acid) microparticles for injectable delivery of vaccine antigens, Adv. Drug Deliv. Rev.,<br />

57(3), 391–410, 2005.<br />

78. Takahira, N., Itoman, M., Higashi, K., Uchiyama, K., Miyabe, M., and Naruse, K., Treatment<br />

outcome of two-stage revision total hip arthroplasty for infected hip arthroplasty using antibioticimpregnated<br />

cement spacer, J. Orthop. Sci., 8(1), 26–31, 2003.<br />

79. Little, S. R., Lynn, D. M., Ge, Q., Anderson, D. G., Puram, S. V., Chen, J., Eisen, H. N., and Langer,<br />

R., Poly-beta amino ester-containing microparticles enhance the activity of nonviral genetic<br />

vaccines, Proc. Natl Acad. Sci. U.S.A., 101(26), 9534–9539, 2004.<br />

80. Lendlein, A., Jiang, H., Junger, O., and Langer, R., Light-induced shape-memory polymers, Nature,<br />

434(7035), 879–882, 2005.<br />

81. Lendlein, A. and Langer, R., Biodegradable, elastic shape-memory polymers for potential biomedical<br />

applications, Science, 296(5573), 1673–1676, 2002.<br />

82. Zion, T. C. and Ying, J. Y., In Glucose responsive nanoparticle for crontrolled insulin delivery.,<br />

Materials Research Society Fall Meeting, Boston, Dec 1–5, 2003.<br />

83. Chan, W. C. and Nie, S., Quantum dot bioconjugates for ultrasensitive nonisotopic detection,<br />

Science, 281(5385), 2016–2018, 1998.<br />

84. Harisinghani, M. G., Barentsz, J., Hahn, P. F., Deserno, W. M., Tabatabaei, S., van de Kaa, C. H., de<br />

la Rosette, J., and Weissleder, R., Noninvasive detection of clinically occult lymph-node metastases<br />

in prostate cancer, N. Engl. J. Med., 348(25), 2491–2499, 2003.<br />

85. Bonnemain, B., Superparamagnetic agents in magnetic resonance imaging: physicochemical characteristics<br />

and clinical applications. A review., J. Drug Target, 6(3), 167–174, 1998.<br />

86. Bala, I., Hariharan, S., and Kumar, M. N., PLGA nanoparticles in drug delivery: the state of the art,<br />

Crit. Rev. Ther. Drug Carrier Syst., 21(5), 387–422, 2004.<br />

87. Drotleff, S., Lungwitz, U., Breunig, M., Dennis, A., Blunk, T., Tessmar, J., and Gopferich, A.,<br />

Biomimetic polymers in pharmaceutical and biomedical sciences, Eur. J. Pharm. Biopharm.,<br />

58(2), 385–407, 2004.<br />

88. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Nanomedicine: current status and future prospects,<br />

Faseb J., 19(3), 311–330, 2005.<br />

89. Panyam, J. and Labhasetwar, V., Biodegradable nanoparticles for drug and gene delivery to cells and<br />

tissue, Adv. Drug Deliv. Rev., 55(3), 329–347, 2003.<br />

90. Ravi Kumar, M., Hellermann, G., Lockey, R. F., and Mohapatra, S. S., Nanoparticle-mediated gene<br />

delivery: state of the art, Expert Opin. Biol. Ther., 4(8), 1213–1224, 2004.<br />

91. Maeda, H., Wu, J., Sawa, T., Matsumura, Y., and Hori, K., Tumor vascular permeability and the EPR<br />

effect in macromolecular therapeutics: a review, J. Control Release, 65(1-2), 271–284, 2000.<br />

92. Liu, D., Mori, A., and Huang, L., Role of liposome size and RES blockade in controlling biodistribution<br />

and tumor uptake of GM1-containing liposomes, Biochim. Biophys. Acta, 1104(1), 95–101, 1992.<br />

93. Alonso, M. J., Gupta, R. K., Min, C., Siber, G. R., and Langer, R., Biodegradable microspheres as<br />

controlled-release tetanus toxoid delivery systems, Vaccine, 12(4), 299–306, 1994.<br />

94. Davda, J. and Labhasetwar, V., Characterization of nanoparticle uptake <strong>by</strong> endothelial cells, Int.<br />

J. Pharm., 233(1-2), 51–59, 2002.<br />

95. Deng, J. S., Li, L., Tian, Y., Ginsburg, E., Widman, M., and Myers, A., In vitro characterization of<br />

polyorthoester microparticles containing bupivacaine, Pharm. Dev. Technol., 8(1), 31–38, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 311<br />

96. Molpeceres, J., Chacon, M., Guzman, M., Berges, L., and del Rosario Aberturas, M., A polycaprolactone<br />

nanoparticle formulation of cyclosporin-A improves the prediction of area under the curve<br />

using a limited sampling strategy, Int. J. Pharm., 187(1), 101–113, 1999.<br />

97. Sommerfeld, P., Sabel, B. A., and Schroeder, U., Long-term stability of PBCA nanoparticle suspensions,<br />

J. Microencapsul., 17(1), 69–79, 2000.<br />

98. Gao, J., Niklason, L., Zhao, X. M., and Langer, R., Surface modification of polyanhydride microspheres,<br />

J. Pharm. Sci., 87(2), 246–248, 1998.<br />

99. Huang, G., Gao, J., Hu, Z., St John, J. V., Ponder, B. C., and Moro, D., Controlled drug release from<br />

hydrogel nanoparticle networks, J. Control Release, 94(2-3), 303–311, 2004.<br />

100. Win, K. Y. and Feng, S. S., Effects of particle size and surface coating on cellular uptake of polymeric<br />

nanoparticles for oral delivery of anticancer drugs, Biomaterials, 26(15), 2713–2722, 2005.<br />

101. Langer, R., Drug delivery. Drugs on target, Science, 293(5527), 58–59, 2001.<br />

102. Langer, R. and Peppas, N. A., Advances in biomaterials, drug delivery, and bionanotechnology,<br />

AICHE J., 49(12), 2990–3006, 2003.<br />

103. Matsusue, Y., Hanafusa, S., Yamamuro, T., Shikinami, Y., and Ikada, Y., Tissue reaction of bioabsorbable<br />

ultra high strength poly (L-lactide) rod. A long-term study in rabbits, Clin. Orthop. Relat.<br />

Res., 317, 246–253, 1995.<br />

104. Chonn, A., Cullis, P. R., and Devine, D. V., The role of surface charge in the activation of the<br />

classical and alternative pathways of complement <strong>by</strong> liposomes, J. Immunol., 146(12), 4234–4241,<br />

1991.<br />

105. Campbell, R. B., Fukumura, D., Brown, E. B., Mazzola, L. M., Izumi, Y., Jain, R. K., Torchilin,<br />

V. P., and Munn, L. L., Cationic charge determines the distribution of liposomes between the<br />

vascular and extravascular compartments of tumors, Cancer Res., 62(23), 6831–6836, 2002.<br />

106. Jaffar, S., Nam, K. T., Khademhosseini, A., Xing, J., Langer, R., and Belcher, A. M., Layer-<strong>by</strong>-layer<br />

surface modification and patterned electrodeposition of quantum dots, Nano Letters, 4(8),<br />

1421–1425, 2004.<br />

107. Wang, Y., Ameer, G. A., Sheppard, B. J., and Langer, R., A tough biodegradable elastomer, Nat.<br />

Biotechnol., 20(6), 602–606, 2002.<br />

108. Farokhzad, O. C., Khademhosseini, A., Jon, S., Hermmann, A., Cheng, J., Chin, C., Kiselyuk, A.,<br />

Teply, B., Eng, G., and Langer, R., Microfluidic system for studying the interaction of nanoparticles<br />

and microparticles with cells, Anal. Chem., 77(17), 5453–5459, 2005.<br />

109. Gref, R., Minamitake, Y., Peracchia, M. T., Domb, A., Trubetskoy, V., Torchilin, V., and Langer, R.,<br />

Poly(ethylene glycol)-coated nanospheres: potential carriers for intravenous drug administration,<br />

Pharm. Biotechnol., 10, 167–198, 1997.<br />

110. Lemarchand, C., Gref, R., and Couvreur, P., Polysaccharide-decorated nanoparticles, Eur. J. Pharm.<br />

Biopharm., 58(2), 327–341, 2004.<br />

111. Lemarchand, C., Gref, R., Passirani, C., Garcion, E., Petri, B., Muller, R., Costantini, D., and<br />

Couvreur, P., Influence of polysaccharide coating on the interactions of nanoparticles with biological<br />

systems, Biomaterials, 27(1), 108–118, <strong>2006</strong>.<br />

112. Weissleder, R., Kelly, K., Sun, E. Y., Shtatland, T., and Josephson, L., Cell-specific targeting of<br />

nanoparticles <strong>by</strong> multivalent attachment of small molecules, Nat. Biotechnol., 23(11), 1418–1423,<br />

2005.<br />

113. Mosqueira, V. C., Legrand, P., Morgat, J. L., Vert, M., Mysiakine, E., Gref, R., Devissaguet, J. P.,<br />

and Barratt, G., Biodistribution of long-circulating PEG-grafted nanocapsules in mice: effects of<br />

PEG chain length and density, Pharm. Res., 18(10), 1411–1419, 2001.<br />

114. Lode, J., Fichtner, I., Kreuter, J., Berndt, A., Diederichs, J. E., and Reszka, R., Influence of surfacemodifying<br />

surfactants on the pharmacokinetic behavior of 14C-poly (methylmethacrylate) nanoparticles<br />

in experimental tumor models, Pharm. Res., 18(11), 1613–1619, 2001.<br />

115. Bazile, D., Prud’homme, C., Bassoullet, M. T., Marlard, M., Spenlehauer, G., and Veillard, M.,<br />

Stealth Me.PEG-PLA nanoparticles avoid uptake <strong>by</strong> the mononuclear phagocytes system, J. Pharm.<br />

Sci., 84(4), 493–498, 1995.<br />

116. McNamara, J. O., Andrechek, E. R., Wang, Y., Viles, K. D., Rempel, R. E., Gilboa, E., Sullenger,<br />

B. A., and Giangrande, P. H., Cell type-specific delivery of siRNAs with aptamer-siRNA chimeras,<br />

Nat. Biotechnol., 24(8), 1005–1015, <strong>2006</strong>.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


312<br />

Nanotechnology for Cancer Therapy<br />

117. Farokhzad, O. C., Karp, J. M., and Langer, R., Nanoparticle-aptamer bioconjugates for cancer<br />

targeting, Expert Opin. Drug Deliv., 3(3), 311–324, <strong>2006</strong>.<br />

118. Chu, T. C., Shieh, F., Lavery, L. A., Levy, M., Richards-Kortum, R., Korgel, B. A., and Ellington,<br />

A. D., Labeling tumor cells with fluorescent nanocrystal-aptamer bioconjugates, Biosens. Bioelectron.,<br />

21(10), 1859–1866, <strong>2006</strong>.<br />

119. Chu, T. C., Marks, J. W., Lavery, L. A., Faulkner, S., Rosenblum, M. G., Ellington, A. D., and Levy,<br />

M., Aptamer:toxin conjugates that specifically target prostate tumor cells, Cancer Res., 66(12),<br />

5989–5992, <strong>2006</strong>.<br />

120. Cheng, J., Teply, B. A., Sherifi, I., Sung, J., Luther, G., Gu, F. X., Levy-Nissenbaum, E., Radovic-<br />

Moreno, A. F., Langer, R., and Farokhzad, O.C., Formulation of functionalized PLGA-PEG nanoparticles<br />

for in vivo targeted drug delivery, Biomaterials, <strong>2006</strong>. In press.<br />

121. Sehgal, D. and Vijay, I. K., A method for the high efficiency of water-soluble carbodiimide-mediated<br />

amidation, Anal. Biochem., 218(1), 87–91, 1994.<br />

122. Takae, S., Akiyama, Y., Otsuka, H., Nakamura, T., Nagasaki, Y., and Kataoka, K., Ligand Density<br />

Effect on Biorecognition <strong>by</strong> PEGylated Gold Nanoparticles: Regulated Interaction of RCA(120)<br />

Lectin with Lactose Installed to the Distal End of Tethered PEG Strands on Gold Surface, Biomacromolecules,<br />

6(2), 818–824, 2005.<br />

123. Polakis, P., Arming antibodies for cancer therapy, Curr. Opin. Pharmacol., 5(4), 382–387, 2005.<br />

124. Waldmann, T. A., Immunotherapy: past, present and future, Nat. Med., 9(3), 269–277, 2003.<br />

125. Wu, A. M. and Senter, P. D., Arming antibodies: prospects and challenges for immunoconjugates,<br />

Nat. Biotechnol., 23(9), 1137–1146, 2005.<br />

126. Bagalkot, V., Farokhzad, O. C., Langer, R., and Jon, S., Aptamer-Doxorubicin physical conjugate as<br />

a novel trageted drug delivery platform, Angew Chem. Int. Ed. Engl., <strong>2006</strong>, In press.<br />

127. Nutiu, R. and Li, Y., Structure-switching signaling aptamers: transducing molecular recognition into<br />

fluorescence signaling, Chemistry, 10(8), 1868–1876, 2004.<br />

128. Srinivasan, J., Cload, S. T., Hamaguchi, N., Kurz, J., Keene, S., Kurz, M., Boomer, R. M. et al., ADPspecific<br />

sensors enable universal assay of protein kinase activity, Chem. Biol., 11(4), 499–508, 2004.<br />

129. Stojanovic, M. N. and Kolpashchikov, D. M., Modular aptameric sensors, J. Am. Chem. Soc.,<br />

126(30), 9266–9270, 2004.<br />

130. Yang, X., Li, X., Prow, T. W., Reece, L. M., Bassett, S. E., Luxon, B. A., Herzog, N. K. et al.,<br />

Immunofluorescence assay and flow-cytometry selection of bead-bound aptamers, Nucleic Acids<br />

Res., 31(10), e54, 2003.<br />

131. Baldrich, E., Restrepo, A., and O’Sullivan, C. K., Aptasensor development: elucidation of critical<br />

parameters for optimal aptamer performance, Anal. Chem., 76(23), 7053–7063, 2004.<br />

132. Yan, X. R., Gao, X. W., Yao, L. H., and Zhang, Z. Q., [Novel methods to detect cytokines <strong>by</strong><br />

enzyme-linked oligonucleotide assay], Sheng Wu Gong Cheng Xue Bao, 20(5), 679–682, 2004.<br />

133. Schou, C. and Heegaard, N. H., Recent applications of affinity interactions in capillary electrophoresis,<br />

Electrophoresis, 27(1), 44–59, <strong>2006</strong>.<br />

134. Dillman, R. O., Johnson, D. E., Shawler, D. L., and Koziol, J. A., Superiority of an acid-labile<br />

daunorubicin-monoclonal antibody immunoconjugate compared to free drug, Cancer Res., 48(21),<br />

6097–6102, 1988.<br />

135. Jaracz, S., Chen, J., Kuznetsova, L. V., and Ojima, I., Recent advances in tumor-targeting anticancer<br />

drug conjugates, Bioorg. Med. Chem., 13(17), 5043–5054, 2005.<br />

136. Hermanson, G. T., Bioconjugate Techniques, Academic Press, San Diego, 1996.<br />

137. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

theory to practice, Pharmacol. Rev., 53(2), 283–318, 2001.<br />

138. Payne, G., Progress in immunoconjugate cancer therapeutics, Cancer Cell, 3(3), 207–212, 2003.<br />

139. Saito, G., Swanson, J. A., and Lee, K. D., Drug delivery strategy utilizing conjugation via reversible<br />

disulfide linkages: role and site of cellular reducing activities, Adv. Drug Deliv. Rev., 55(2), 199–215, 2003.<br />

140. Trail, P. A., Willner, D., Bianchi, A. B., Henderson, A. J., TrailSmith, M. D., Girit, E., Lasch, S.,<br />

Hellstrom, I., and Hellstrom, K. E., Enhanced antitumor activity of paclitaxel in combination with<br />

the anticarcinoma immunoconjugate BR96-doxorubicin, Clin. Cancer Res., 5(11), 3632–3638, 1999.<br />

141. Mosure, K. W., Henderson, A. J., Klunk, L. J., and Knipe, J. O., Disposition of conjugate-bound and<br />

free doxorubicin in tumor-bearing mice following administration of a BR96-doxorubicin immunoconjugate<br />

(BMS 182248), Cancer Chemother. Pharmacol., 40(3), 251–258, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Aptamers and Cancer Nanotechnology 313<br />

142. Lu, P. Y., Xie, F. Y., and Woodle, M. C., siRNA-mediated antitumorigenesis for drug target<br />

validation and therapeutics, Curr. Opin. Mol. Ther., 5(3), 225–234, 2003.<br />

143. Cheng, J. C., Moore, T. B., and Sakamoto, K. M., RNA interference and human disease, Mol. Genet.<br />

Metab., 80(1-2), 121–128, 2003.<br />

144. Cheng, J. C. and Sakamoto, K. M., The emerging role of RNA interference in the design of novel<br />

therapeutics in oncology, Cell Cycle, 3(11), 1398–1401, 2004.<br />

145. Smith, A. M., Gao, X., and Nie, S., Quantum dot nanocrystals for in vivo molecular and cellular<br />

imaging, Photochem. Photobiol., 80(3), 377–385, 2004.<br />

146. Gao, X., Cui, Y., Levenson, R. M., Chung, L. W., and Nie, S., In vivo cancer targeting and imaging<br />

with semiconductor quantum dots, Nat. Biotechnol., 22(8), 969–976, 2004.<br />

147. Gao, X., Yang, L., Petros, J. A., Marshall, F. F., Simons, J. W., and Nie, S., In vivo molecular and<br />

cellular imaging with quantum dots, Curr. Opin. Biotechnol., 16(1), 63–72, 2005.<br />

148. Hayat, M., Colloidal Gold: Principles, Methods and Applications, Academic Press, San Diego,<br />

1989.<br />

149. West, J. L. and Halas, N. J., Engineered nanomaterials for biophotonics applications: improving<br />

sensing, imaging, and therapeutics, Annu. Rev. Biomed. Eng., 5, 285–292, 2003.<br />

150. Loo, C., Hirsch, L., Lee, M. H., Chang, E., West, J., Halas, N., and Drezek, R., Gold nanoshell<br />

bioconjugates for molecular imaging in living cells, Opt. Lett., 30(9), 1012–1014, 2005.<br />

151. Loo, C., Lowery, A., Halas, N., West, J., and Drezek, R., Immunotargeted nanoshells for integrated<br />

cancer imaging and therapy, Nano Lett., 5(4), 709–711, 2005.<br />

152. Xu, S., Ji, X., Xu, W., Li, X., Wang, L., Bai, Y., Zhao, B., and Ozaki, Y., Immunoassay using probelabelling<br />

immunogold nanoparticles with silver staining enhancement via surface-enhanced Raman<br />

scattering, Analyst, 129(1), 63–68, 2004.<br />

153. Driskell, J. D., Kwarta, K. M., Lipert, R. J., Porter, M. D., Neill, J. D., and Ridpath, J. F., Low-level<br />

detection of viral pathogens <strong>by</strong> a surface-enhanced Raman scattering based immunoassay, Anal.<br />

Chem., 77(19), 6147–6154, 2005.<br />

154. Copland, J. A., Eghtedari, M., Popov, V. L., Kotov, N., Mamedova, N., Motamedi, M., and<br />

Oraevsky, A. A., Bioconjugated gold nanoparticles as a molecular based contrast agent: implications<br />

for imaging of deep tumors using optoacoustic tomography, Mol. Imaging Biol., 6(5), 341–349,<br />

2004.<br />

155. Levy, M., Cater, S. F., and Ellington, A. D., Quantum-dot aptamer beacons for the detection of<br />

proteins, Chem. biochem., 6(12), 2163–2166, 2005.<br />

156. Meyer, T. and Teruel, M. N., Fluorescence imaging of signaling networks, Trends Cell Biol., 13(2),<br />

101–106, 2003.<br />

157. Mank, M., Reiff, D. F., Heim, N., Friedrich, M. W., Borst, A., and Griesbeck, O., A FRET-Based<br />

Calcium Biosensor with Fast Signal Kinetics and High Fluorescence Change, Biophys. J., 90(5),<br />

1790–1796, <strong>2006</strong>.<br />

158. Chan, W. C., Maxwell, D. J., Gao, X., Bailey, R. E., Han, M., and Nie, S., Luminescent quantum dots<br />

for multiplexed biological detection and imaging, Curr. Opin. Biotechnol., 13(1), 40–46, 2002.<br />

159. Han, M., Gao, X., Su, J. Z., and Nie, S., Quantum-dot-tagged microbeads for multiplexed optical<br />

coding of biomolecules, Nat. Biotechnol., 19(7), 631–635, 2001.<br />

160. Huang, C. C., Huang, Y. F., Cao, Z., Tan, W., and Chang, H. T., Aptamer-modified gold nanoparticles<br />

for colorimetric determination of platelet-derived growth factors and their receptors, Anal.<br />

Chem., 77(17), 5735–5741, 2005.<br />

161. Elghanian, R., Storhoff, J. J., Mucic, R. C., Letsinger, R. L., and Mirkin, C. A., Selective colorimetric<br />

detection of polynucleotides based on the distance-dependent optical properties of gold nanoparticles,<br />

Science, 277(5329), 1078–1081, 1997.<br />

162. Storhoff, J. J., Marla, S. S., Bao, P., Hagenow, S., Mehta, H., Lucas, A., Garimella, V. et al., Gold<br />

nanoparticle-based detection of genomic DNA targets on microarrays using a novel optical detection<br />

system, Biosens. Bioelectron., 19(8), 875–883, 2004.<br />

163. Storhoff, J. J., Lucas, A. D., Garimella, V., Bao, Y. P., and Muller, U. R., Homogeneous detection of<br />

unamplified genomic DNA sequences based on colorimetric scatter of gold nanoparticle probes, Nat.<br />

Biotechnol., 22(7), 883–887, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


17<br />

CONTENTS<br />

Polymeric Micelles for<br />

Formulation of Anti-Cancer<br />

Drugs<br />

Helen Lee, Patrick Lim Soo, Jubo Liu,<br />

Mark Butler, and Christine Allen<br />

17.1 Introduction ........................................................................................................................318<br />

17.2 General Properties of Polymeric Micelles for Drug Formulation ....................................319<br />

17.2.1 Building Blocks for Polymeric Micelles .............................................................319<br />

17.2.2 Physico-Chemical Properties of Polymeric Micelles ..........................................321<br />

17.2.2.1 Size and Size Distribution ..................................................................321<br />

17.2.2.2 Morphology .........................................................................................321<br />

17.2.2.3 Stability ...............................................................................................322<br />

17.2.3 Preparation of Block Copolymer Micelle Formulations .....................................323<br />

17.2.4 Properties of Micelle Formulations .....................................................................324<br />

17.2.4.1 Drug Partitioning.................................................................................324<br />

17.2.4.2 Factors Affecting Drug Loading .........................................................324<br />

17.2.4.3 Factors Influencing Drug Release from Polymeric Micelles .............325<br />

17.3 Physical Characteristics of Polymeric Micelles for Clinical<br />

Applications in Cancer Therapy ........................................................................................326<br />

17.3.1 Doxorubicin ..........................................................................................................327<br />

17.3.1.1 NK911 .................................................................................................327<br />

17.3.1.2 SP1049C ..............................................................................................328<br />

17.3.2 Paclitaxel ..............................................................................................................329<br />

17.3.2.1 Genexol-PM ........................................................................................329<br />

17.3.2.2 NK105 .................................................................................................330<br />

17.4 Interactions of Polymeric Micelles with Cancer Cells......................................................330<br />

17.4.1 Drug Diffusion......................................................................................................331<br />

17.4.2 Cellular Internalization.........................................................................................332<br />

17.4.3 In Vitro Cytotoxicity ............................................................................................334<br />

17.4.4 Influence or Effect on Multi-Drug Resistance.....................................................335<br />

17.5 In Vivo Biological Performance of Block Copolymer Micelles ......................................336<br />

17.5.1 Fate of Polymeric Micelles In Vivo ....................................................................337<br />

17.5.2 In Vivo Drug Release Profiles .............................................................................338<br />

17.5.3 Toxicity, Pharmacokinetics, and Anti-Cancer Efficacy of Drugs<br />

Formulated in Block Copolymer Micelles ..........................................................339<br />

17.5.3.1 Doxorubicin Formulations (SP1049C and NK911) ...........................339<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

317


318<br />

17.5.3.2 Paclitaxel Formulations (Genexol-PM and NK105) ..........................339<br />

17.6 Evolution of Polymeric Micelles .......................................................................................340<br />

17.6.1 Triggered Stimulation of Drug Activation or Release ........................................340<br />

17.6.1.1 Ultrasonic Irradiation ..........................................................................340<br />

17.6.1.2 Thermosensitive Systems ....................................................................341<br />

17.6.1.3 Photodynamic Therapy .......................................................................342<br />

17.6.2 Active Targeting...................................................................................................342<br />

17.6.2.1 Ligand Coupling..................................................................................343<br />

17.6.2.2 pH-Responsive Systems ......................................................................344<br />

17.7 Future Perspectives.............................................................................................................345<br />

Acknowledgments .........................................................................................................................345<br />

References .....................................................................................................................................345<br />

17.1 INTRODUCTION<br />

Nanotechnology for Cancer Therapy<br />

The use of conventional formulations for the systemic administration of many small molecule anticancer<br />

agents often results in an unsatisfactory therapeutic effect because of the narrow therapeutic<br />

window of these drugs. The low therapeutic index of these drugs is mostly attributed to the doselimiting<br />

toxicities that are associated with them and/or the excipients used in the formulation.<br />

Therefore, colloidal carriers that are composed of biocompatible and biodegradable materials<br />

with sustained drug release have been developed for formulation and delivery. Nanosized colloidal<br />

carriers have been shown to allow for increased accumulation of drug at tumors via the enhanced<br />

permeation and retention (EPR) effect. 1 As a result of the altered pharmacokinetics, the exposure of<br />

chemotherapeutic drugs to healthy tissues is reduced. Colloidal carriers including liposomes,<br />

nanoparticles, and polymeric micelles have been explored extensively for the delivery of a wide<br />

range of drugs as anti-cancer therapy. 1,2<br />

Polymeric micelles are nanosized assemblies of amphiphilic block copolymers that are suitable<br />

for the delivery of hydrophobic and amphiphilic agents. In an aqueous medium, micelles consist of<br />

a hydrophilic shell that minimizes clearance <strong>by</strong> the mononuclear phagocytic system (MPS) and a<br />

hydrophobic core that functions as a reservoir for hydrophobic drugs. In the past two decades, four<br />

polymeric micelle formulations loaded with chemotherapeutic drugs (i.e., NK911, SP1049C,<br />

Genexol-PM, and NK105) have entered clinical trial development. The results from the clinical<br />

studies have indicated that the polymeric micelle formulations reduce the toxicity associated with<br />

conventional formulations of these drugs that, in turn, results in a higher therapeutic index. Many<br />

preclinical fundamental studies have evaluated the relationships between the composition of the<br />

copolymers and the physico-chemical properties of the micelles. The properties of the micelles<br />

such as polymer–drug compatibility, thermodynamic and kinetic stability, and the drug release<br />

profiles have been shown to influence the in vivo performance and therapeutic effectiveness of the<br />

micelle-formulated drugs. These studies serve as guidelines for the optimization of polymeric<br />

micelles for clinical applications.<br />

This chapter is divided into six major sections and provides a discussion of the various aspects<br />

relating to use of polymeric micelles for formulation of anti-cancer drugs. Section 17.2 contains<br />

information on the physico-chemical properties of micelles (e.g., composition, size and size distribution,<br />

morphology, stability), micelle preparation techniques as well as drug loading and release<br />

properties of micelles. Section 17.3 highlights the preclinical development of several polymeric<br />

micelle formulations that are currently under clinical trial evaluation for use as cancer therapy.<br />

Specifically, the optimization of the physico-chemical characteristics of NK911, SP1049C,<br />

Genexol-PM, and NK105 are discussed. In Section 17.4, the interaction between polymeric<br />

micelles and cancer cells, including cellular internalization, in vitro cytotoxicity, and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 319<br />

chemosensitization of multi-drug resistant (MDR) cancer cells is reviewed. In Section 17.5, the<br />

physico-chemical properties of polymeric micelles are related to their in vivo performance such as<br />

in vivo drug release, pharmacokinetics, toxicity profiles, and anti-cancer efficacy. Finally, in<br />

Section 17.6, the development of several advanced polymeric micelle systems that are capable<br />

of targeted delivery to tumors via external stimuli (e.g., ultrasound, heat, light) or active targeting<br />

mechanisms (e.g., ligand coupling, pH-sensitive) are introduced.<br />

17.2 GENERAL PROPERTIES OF POLYMERIC MICELLES FOR DRUG<br />

FORMULATION<br />

In order to fully exploit the benefits of polymeric micelles for drug delivery, the drug-loaded<br />

micelle system requires extensive optimization in terms of its physico-chemical characteristics.<br />

Factors that may influence the performance of drug-loaded micelle systems include the selection of<br />

the building blocks to form the micelles, physico-chemical properties, drug partitioning, drug<br />

loading, and drug release profile.<br />

17.2.1 BUILDING BLOCKS FOR POLYMERIC MICELLES<br />

Amphiphilic block copolymers are the typical building blocks used for the formation of micelles.<br />

The term amphiphilic refers to a molecule that consists of both hydrophobic and hydrophilic<br />

segments. In aqueous solution, these block copolymers can self-assemble to form micelles<br />

consisting of a hydrophobic core and a hydrophilic corona or shell. 3–6 The hydrophilic corona<br />

(a)<br />

Solvent<br />

(b)<br />

Solvent<br />

(c)<br />

+ Water<br />

or buffer<br />

Remove<br />

Solvent<br />

+ Water<br />

or buffer<br />

+<br />

Drug<br />

Remove<br />

Solvent<br />

+ Water<br />

or buffer<br />

or<br />

Hydrophilic corona<br />

Hydrophobic core<br />

FIGURE 17.1 Techniques for preparation of drug-loaded block copolymer micelles. (a) Copolymers and<br />

drugs are directly dissolved in water and form drug-loaded micelles spontaneously using the direct dissolution<br />

method. (b) In the evaporation method, copolymers and drugs are dissolved in an organic solvent. The solvent<br />

is evaporated, and drug-loaded micelles are formed with the addition of an aqueous solution. (c) In the dialysis<br />

method, copolymers and drugs are dissolved in an organic solvent that is miscible with water. Subsequent<br />

addition of water induces formation of micelles with swollen cores. Finally, the solvent is removed <strong>by</strong><br />

evaporation or dialysis to form intact drug-loaded micelles. The structure of a drug-loaded micelle is also<br />

shown above.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


320<br />

TABLE 17.1<br />

Examples of Polymers That Are Commonly Used as Building Blocks for Block Copolymer<br />

Micelles in Drug Delivery<br />

Hydrophilic segment (corona-forming block)<br />

Poly(ethylene glycol) or<br />

poly(ethylene oxide)<br />

PEG or PEO CH2 CH2 O<br />

m<br />

Poly(N-vinyl-pyrrolidone) PVP CH2 CH<br />

N O<br />

m<br />

Hydrophobic segment (core-forming block)<br />

Poly(caprolactone) PCL<br />

Poly(D,L-lactide) PDLLA<br />

O<br />

C CH<br />

Poly(glycolide) PGA<br />

O<br />

C (CH2) 5 O<br />

n<br />

acts as the interface between the core and the exterior environment, and the hydrophobic core<br />

serves as the reservoir for the incorporation of lipophilic drugs (Figure 17.1).<br />

In drug delivery, biodegradable and biocompatible copolymers are chosen as a result of their<br />

degradation into non-toxic oligomers or monomers that are eventually absorbed in the body and/or<br />

eliminated. Many of the amphiphilic block copolymers used in drug delivery contain either a<br />

polyester, polyether, or a poly(amino acid) derivative as the hydrophobic block. A number of<br />

these hydrophobic polymers that have been used as the core of a micelle are listed in Table 17.1.<br />

Polyesters such as poly(D,L-lactide) (PDLLA), poly(glycolide) (PGA), and poly(3-caprolactone)<br />

(PCL) have been chosen as the core block because these polymers are biocompatible, biodegradable,<br />

and FDA approved for use in medical devices. Poly(propylene oxide) (PPO) is a polyether that<br />

has been used as the hydrophobic block in the well-studied Pluronic w family, the triblock<br />

CH 3<br />

O<br />

n<br />

O<br />

C CH2 O<br />

n<br />

O<br />

O<br />

Poly(D,L-lactide-co-glycolide) PLGA C CH2 O<br />

x<br />

C CH O<br />

y<br />

O<br />

O<br />

Poly(aspartic acid) PAsp NH C CH NH<br />

CH2 COOH<br />

x<br />

C CH2 CH NH<br />

COOH<br />

y<br />

Poly(b-benzyl-L-aspartate) PBLA NH<br />

O<br />

C CH NH<br />

n<br />

Pluronic CH2 COOC H2 w copolymer (triblock)<br />

O CH2 CH2 O CH CH2 n<br />

O CH 2 CH 2<br />

m/2<br />

m/2<br />

CH3 Poly(ethylene oxide)-b-poly(propylene oxide)-b-poly(ethylene oxide)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

CH 3


Polymeric Micelles for Formulation of Anti-Cancer Drugs 321<br />

copolymers of poly(ethylene glycol)-block-poly(propylene oxide)-block-poly(ethylene glycol)<br />

(PEG-b-PPO-b-PEG). 7 Finally, poly(amino acid)s such as poly(aspartic acid) (PAsp) and poly<br />

(b-benzyl-L-aspartate) (PBLA) have also been widely explored as materials for the hydrophobic<br />

block because they can be easily modified synthetically. The wide range of materials available for<br />

selection as the hydrophobic block of the copolymer enables the preparation of micelles’ having a<br />

variety of microenvironments within their cores.<br />

In contrast, the hydrophilic block is important for stabilizing the micelle in its aqueous environment.<br />

The criteria for the selection of the hydrophilic block is that the polymer should be both<br />

uncharged and water soluble (e.g., poly(ethylene glycol), polyacrylamide, polyhydroxyethylmethacrylate,<br />

poly(N-vinyl-pyrrolidone) (PVP), and poly(vinyl alcohol)). 8 Poly(ethylene glycol) (PEG)<br />

is most commonly used as the hydrophilic segment in micelles. The molecular weights of PEG used<br />

in block copolymers typically range between 1000 and 12,000 g/mol with a chain length that is<br />

typically equal to or greater than the length of the core-forming block. 4 PEG can also be referred to<br />

as poly(ethylene oxide) (PEO), a polymer with a molecular weight that is greater than 20,000 g/<br />

mol, whereas PEG refers to polymers with molecular weights below this value. 9 Because of its high<br />

water solubility and large excluded volume, PEG can stabilize the micelles <strong>by</strong> sterically excluding<br />

other polymers from the surface of the micelles. 1,2,10 In addition, PEG can inhibit the surface<br />

adsorption of biological components such as proteins, improve the residence time of the micelles<br />

in the circulation, and limit the micelle clearance <strong>by</strong> the MPS. 10–13<br />

The selection of building blocks for the micelle system includes the type of polymers and the<br />

length of the core- and corona-forming polymer blocks; these will influence the physico-chemical<br />

properties of the micelles and also affect the overall therapeutic efficacy of the delivery system.<br />

17.2.2 PHYSICO-CHEMICAL PROPERTIES OF POLYMERIC MICELLES<br />

In order to design an effective micellar drug delivery system, several key physico-chemical properties<br />

of the micelles should be considered as a means to optimize performance. These include<br />

micelle size and size distribution, morphology, and stability.<br />

17.2.2.1 Size and Size Distribution<br />

Block copolymer micelles typically range in size from 10 to 100 nm. As a result of their small size<br />

and the steric stabilization provided <strong>by</strong> the hydrophilic corona, micelles are less susceptible to MPS<br />

clearance. 14,15 In addition, their small size allows them to extravasate through the leaky endothelium<br />

and accumulate at the tumor interstitium. Therefore, micelles can passively target tumors<br />

via the enhanced permeability and retention (EPR) effect. 1,16 Micelles can also pass through the<br />

pores of the renal filtration system as individual polymer chains following disassembly. A renal<br />

threshold limit of less than 50,000 g/mol has been reported for polymeric carriers, 17 so it is<br />

important to design a micelle system where the individual polymer chains have a molecular<br />

weight that is lower than this threshold limit. The small size of the micelles also allows for<br />

sterilization <strong>by</strong> filtration in order to remove bacteria prior to intravenous administration.<br />

Filtration can also be used to reduce the size distribution of colloidal carriers that have broad<br />

distributions following initial preparation. Micelles tend to have a relatively narrow size distribution;<br />

however, they are prone to secondary aggregation because of insufficient coverage of the<br />

hydrophobic core <strong>by</strong> the hydrophilic chains. Dilution has been shown to be useful in breaking apart<br />

the secondary aggregates of primary micelles. 18<br />

17.2.2.2 Morphology<br />

When the corona-forming block of a copolymer is smaller than the core-forming block, crew-cut<br />

micelles are produced. 3 These copolymers may be used to prepare a wide range of morphologies<br />

including spheres, rods, lamellae, vesicles, large compound micelles, etc. 19,20 Of particular interest<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


322<br />

are block copolymer vesicles that may prove to be effective for the simultaneous delivery of both<br />

hydrophilic and hydrophobic agents. 21,22 However, most of the block copolymer micelles used in<br />

drug delivery consist of spherical, star-type micelles as they are formed from copolymers that have<br />

a hydrophilic block that is of equal or greater length than the hydrophobic core-forming block. 3 A<br />

sample transmission electron micrograph of a population of spherical micelles is shown in<br />

Figure 17.2.<br />

17.2.2.3 Stability<br />

ES1 B, 100k.tif<br />

PEO–BTMC micelle<br />

PEO–BTMC micelle<br />

Print mag: 101000× @ 112 mm<br />

16:30 01/14/04<br />

Nanotechnology for Cancer Therapy<br />

100 nm<br />

HV=75 kV<br />

Direct mag: 100000×<br />

FIGURE 17.2 Transmission electron micrograph of block copolymer micelles formed from methoxy poly<br />

(ethylene glycol)-block-poly(5-benxyloxy-trimethylene carbonate) copolymers. (Reproduced with permission<br />

from Biomacromolecules 2004, 5, 1810–1817. Copyright 2004, American Chemical Society).<br />

The physical stability of block copolymer micelles is important because the early release of the<br />

drug into the bloodstream as a result of the disassembly of copolymer chains results in insufficient<br />

accumulation of the drug at the target site. It is critical that the copolymer micelles possess<br />

sufficient stability to remain intact in the circulation. The thermodynamic tendency for micelles<br />

to disassemble into individual chains is reflected <strong>by</strong> the critical micelle concentration (CMC) of the<br />

copolymer. Below the CMC, the copolymer is in the form of single polymer chains or unimers, and<br />

above the CMC, the copolymer exists as micelles in equilibrium with a small population of single<br />

chains. In this way, the CMC represents an important parameter as it determines the thermodynamic<br />

stability of the micelles during dilution. 23 Small molecule surfactants form micelles<br />

that tend to have higher CMC values that make them more susceptible to disassembly following<br />

dilution. In contrast, amphiphilic block copolymers tend to be more thermodynamically stable with<br />

lower CMC values in the range of 10 K7 –10 K6 M. 3,24,25<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 323<br />

The CMC has been found to be dependent on the nature and length of the hydrophobic block,<br />

the length of the hydrophilic block, and the total molecular weight of the copolymer. 26,27 Specifically,<br />

increasing the hydrophobicity of the core-forming block reduces the CMC and, in turn,<br />

improves the thermodynamic stability of the micelles. For example, Leroux et al. have demonstrated<br />

that increasing the hydrophobic PDLLA block from 27 to 55 mol% in a series of PVP-b-<br />

PDLLA-b-PVP copolymers results in a decrease in the CMC from 19.9 to 5.1 mg/L. 28 In contrast,<br />

when the hydrophilic block length is increased and the hydrophobic block is kept constant, the<br />

CMC increases as has been shown for the Pluronic w copolymers. 29<br />

Even at concentrations below the CMC, micelles can remain kinetically stable for extended<br />

periods of time depending on their glass transition temperature (Tg) and/or Tm if the core-forming<br />

block is semicrystalline. Micelles with a core-forming block that has a lower Tg (below 378C) will<br />

tend to disassemble at a faster rate than those with a higher Tg as a result of the free motion of the<br />

core-forming polymer chains at temperatures above the Tg. For example, Kataoka et al. found a<br />

dramatic increase in the CMC of PEG-b-PDLLA micelles at a temperature above the Tg of the<br />

hydrophobic PDLLA polymer that suggested a decrease in the kinetic stability of the system. 30<br />

Kang et al. improved the kinetic stability of PEG-b-poly(lactide) (PLA) block copolymer micelles<br />

<strong>by</strong> blending equimolar mixtures of PEG-b-poly(D-lactide) (PDLA) and PEG-b-poly(L-lactide)<br />

(PLLA) enantiomeric copolymers. They reported superior kinetic stability as a result of the<br />

increased van der Waals interactions between the PLA polymer chains that promoted a denser<br />

packing and a tighter conformation. 31 Finally, the presence of the hydrophobic drug in the core has<br />

also been shown to promote kinetic stability of the formulation. 32<br />

17.2.3 PREPARATION OF BLOCK COPOLYMER MICELLE FORMULATIONS<br />

The selection of the appropriate preparative technique for micelle formation is largely based on<br />

the solubility of the copolymer and the drug in aqueous media. There are several methods<br />

that have been developed for the preparation of drug-loaded block copolymer micelles,<br />

including the direct dissolution method, evaporation method, and dialysis method as depicted<br />

in Figure 17.1.<br />

The direct dissolution method is employed for block copolymers that are relatively soluble in<br />

water (e.g., Pluronic w ). In this method, the copolymer and the drug are directly dissolved in water<br />

or an aqueous solution such as phosphate-buffered saline (PBS) with subsequent heating and/or<br />

stirring to induce micellization. Successful loading of doxorubicin, etoposide, nystatin, and haloperidol<br />

into Pluronic w micelles has been achieved using this method. 33,34<br />

For copolymers and drugs with limited aqueous solubility, the evaporation method can be used<br />

to prepare the micelle formulations. This method involves first dissolving the copolymer and the<br />

drug in a common solvent or a mixture of two miscible solvents. The mixture is then stirred, and the<br />

solvent is allowed to evaporate, yielding a copolymer–drug film that can be reconstituted in warm<br />

water or a buffer to form the drug-loaded micelles. 35–38<br />

The most commonly used method for the preparation of drug-loaded micelles from poorly<br />

water-soluble copolymers is the dialysis method. The copolymer and the drug are first dissolved in a<br />

common organic solvent that is miscible with water. Slow addition of water results in the formation<br />

of micelles with swollen cores. The solution is then dialyzed against a large volume of water to<br />

remove the organic solvent. 3,28,32,39 As the percentage of solvent decreases, the degree of swelling<br />

in the micelle core decreases to form intact drug-loaded micelles. Alternatively, the solvent can be<br />

removed <strong>by</strong> evaporation of the solvent from the polymer–drug mixture to form micelles (i.e.,<br />

emulsification method). 38,40,41 The selection of the organic solvent and the ratio of water-to-solvent<br />

employed have been shown to greatly affect the physical properties and drug loading of the<br />

micelles. 39<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


324<br />

17.2.4 PROPERTIES OF MICELLE FORMULATIONS<br />

The drug loading and drug release profiles of block copolymer micelle formulations will influence<br />

the in vivo therapeutic efficacy of the system. There is a balance between the amount of drug that<br />

can be loaded into the hydrophobic core and the amount of drug that can be released from the same<br />

core. 36 The drug loading and drug release from the micelles is highly dependent on the partitioning<br />

behavior of the drug into the micelle.<br />

17.2.4.1 Drug Partitioning<br />

Drug loading into block copolymer micelles is achieved <strong>by</strong> the partitioning of the lipophilic agents<br />

into the micelle core. The partition coefficient (Kv) refers to the extent to which a drug selectively<br />

partitions into a micelle and can be expressed as follows:<br />

K v Z ½DrugŠ micelle<br />

½DrugŠ aqueous<br />

(17.1)<br />

where [Drug] micelle and [Drug] aqueous represent the molar concentrations of the drug in the micelle<br />

phase and aqueous phase, respectively. 23,42 Many different experimental methods including surface<br />

tension, 23 UV–vis spectroscopy, 43,44 fluorescence, 33,42,45,46 solubility in model solvents, 47,48 and<br />

theoretical methods 49,50 have been used to study the partitioning behavior of molecules in block<br />

copolymer micelle solutions. The length of the hydrophobic block influences both the CMC and the<br />

partition coefficient. An increase in either the length or the hydrophobicity of the core-block has<br />

been shown to improve the partitioning behavior of hydrophobic agents into the micelle core. 51,52<br />

Hydrophobic molecules such as pyrene preferentially reside in the hydrophobic micelle core. 50<br />

However, the micelle core is not the only site for drug solubilization. Various amphiphilic solubilizates<br />

can reside at the core–corona interface or in the corona itself. 45,47,53<br />

17.2.4.2 Factors Affecting Drug Loading<br />

Nanotechnology for Cancer Therapy<br />

The partitioning behavior of drugs into the micelle core, core–corona interface and/or corona<br />

controls the amount of drug that can be solubilized in the micelles. However, for a block copolymer<br />

micelle system to act as a true carrier, it must retain the drug within the micelles for prolonged<br />

periods of time following administration. Also, it is highly unlikely that a single block copolymer<br />

micelle system can effectively deliver all types of drugs. Therefore, selecting a core block for the<br />

drug of interest is an important consideration in the design of an effective micelle delivery system.<br />

Micelle drug loading is influenced <strong>by</strong> several factors, including the hydrophobicity and length of<br />

the core-forming block, the length of the corona-forming block, the specific interactions between<br />

the drug and core, and the method of micelle preparation.<br />

The hydrophobicity and length of the core-forming block have a great influence on the size of<br />

the micelles and the loading into the core of the micelle. For example, Lin et al. reported that<br />

increasing the degree of hydrophobicity of the core-forming block improved the drug loading into<br />

the micelles. 54 Also a larger amount of drug can be incorporated into micelles formed from<br />

copolymers having a larger core-forming block in comparison to a shorter hydrophobic block.<br />

For example, Shuai et al. demonstrated that for doxorubicin-loaded micelles formed from PEGb-PCL<br />

copolymers with a constant PEG block length, increasing the PCL block length resulted in<br />

an increase in doxorubicin loading from 3.1% to 4.3% (wt./wt.). The hydrophilic block length can<br />

also influence drug loading. By increasing the corona-forming block, the CMC will also increase,<br />

and this may result in a decrease in the aggregation number (N agg). For example, Gadelle et al.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 325<br />

showed that increasing the PEG block length caused an increase in the CMC, a decrease in the Nagg,<br />

and a decrease in the degree of solubilization of hydrophobic agent into the micelles. 55<br />

Specific interactions between the core-forming block and entrapped drug can also affect drug<br />

loading in the micelles. This is referred to as polymer–drug compatibility where the compatibility is<br />

defined as interaction between the two species that does not involve any chemical change to either<br />

the drug or the polymer. Typically, the interactions that are present between polymer and drug are<br />

hydrogen bonding, ionic, and pi–pi. Lee et al. showed that <strong>by</strong> increasing the number of functionalized<br />

carboxylic acid (COOH) groups from 0% to 19.5% (weight percent of repeat units), they were<br />

able to augment papaverine loading from 3.5% to 15% (wt./wt.) in PEG-b-PDLLA micelles. They<br />

attributed the increase in papaverine loading levels to hydrogen bonding interactions between the<br />

COOH groups of the copolymer and the unpaired electron groups of the oxygen and/or nitrogen<br />

atoms in the drug. 56 Other approaches include improving drug loading <strong>by</strong> enhancing the local<br />

microenvironment of the micelle core through the means of conjugation of a small amount of<br />

drug 16 or conjugation of side chains with similar structures to the drug. 57,58<br />

Finally, the method that is selected for preparation of the micelles has also been shown to have<br />

an effect on the overall degree of drug loading. Yokoyama et al. have shown that using different<br />

micelle preparation techniques led to distinctly different drug loading efficiencies for<br />

campthothecin (CPT) in a series of PEG-b-poly(aspartate ester) copolymers. 38,59 The use of the<br />

dialysis method resulted in a CPT loading efficiency of 19% and a cloudy formulation with large<br />

aggregates and drug precipitates. In contrast, when CPT was loaded using the evaporation method,<br />

a loading efficiency of 73% was obtained and also a clear solution was observed. 38 In addition to the<br />

selection of the preparation method, method-specific parameters (nature of organic solvent, solvent<br />

ratio) have also been proven to affect drug loading into the micelles. Yokoyama et al. showed that<br />

the use of dimethylformamide to prepare KRN 5500 loaded PEG-b-poly(16-b-cetyl-L-aspartate-cob-benzyl-L-aspartate)<br />

micelles resulted in a higher drug loading level when compared to using<br />

dimethylsulfoxide. 60 The ability to achieve a high drug loading level is important as this allows for<br />

maximization of the drug to polymer ratio that, in turn, means that less copolymer will need to be<br />

administered per dose of formulation.<br />

17.2.4.3 Factors Influencing Drug Release from Polymeric Micelles<br />

A drug can be localized in potentially three different areas in a block copolymer micelle: in the core,<br />

at the core–corona interface, and at the corona. 53 The drug in the micelle core has a longer diffusion<br />

path relative to the drug that is located at the core–corona interface or the corona. The drug that is<br />

released from the core–corona interface or the corona is typically referred to as a burst release, an<br />

initial discharge of drug in a relatively short period of time (typically hours). For example, Bromberg<br />

et al. observed a burst release (5–11%) of b-estradiol from the poly(acrylic acid) (PAA) corona<br />

of Pluronic w -PAA micelles. 61 In most cases, after a burst release, a slower and more consistent<br />

release that typically lasts for a longer period of time (typically days to months) is observed. For<br />

example, Kim et al. demonstrated slow release of indomethacin (i.e., less than 30% released over a<br />

period of approximately 100 h) from Pluronic w /PCL micelles. 62<br />

The kinetics of drug release from micellar systems has been studied extensively <strong>by</strong> many<br />

groups. 36,45,61,63–65 Despite the extensive release studies, the mechanism of drug release from<br />

block copolymer micelles is not fully understood. However, many groups believe that diffusion<br />

and copolymer degradation are the two principal mechanisms of release provided that the micelles<br />

are stable under the release conditions investigated. 63 If the interaction between the polymer and the<br />

drug is strong and the rate of biodegradation is fast, then the degradation rate would govern the rate<br />

of drug release. However, the rate of drug release from micelles, as observed in vitro, usually<br />

exceeds the rate of copolymer degradation. As a result, polymer degradation can most often be<br />

ruled out as one of the main mechanisms of drug release. Therefore, diffusion of the drug may be<br />

considered the principal mechanism of release from micelles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


326<br />

Nanotechnology for Cancer Therapy<br />

Drug release is influenced <strong>by</strong> three main factors: the characteristics of the drug (amount of drug,<br />

molecular weight of the drug, and molecular volume of the drug), the properties of the core-forming<br />

block (molecular weight of the core, nature of the core, and physical state of the core), and the<br />

degree of polymer–drug compatibility.<br />

The amount of drug present in the micelle has been found to influence the release. Most often it<br />

has been found that an increase in the concentration of drug present results in a decrease in the rate<br />

of drug release. 36,65–67 For example, Gref et al. reported that higher concentrations of lidocaine in<br />

PEG-b-poly(D,L-lactide-co-glycolide) (PEG-b-PLGA) micelles resulted in a slower release rate for<br />

the drug. 66 Similarly, Lim Soo et al. observed that increasing the amount of estradiol in PEG-b-PCL<br />

micelles decreased the rate of drug release. 65<br />

The physical properties of the drug, including molecular weight and molecular volume have<br />

also been shown to influence their rate of release from micelles. An increase in the molecular<br />

weight of the drug forces a greater reorientation of the polymer chains for movement of the drug<br />

molecules through the polymer matrix. An increase in the volume of the drug will also require a<br />

greater effort to reorganize the polymer to accommodate its movement. As a result, larger or<br />

heavier molecular weight drugs will tend to have lower diffusion coefficients as opposed to<br />

smaller molecular weight agents.<br />

The rate of diffusion of a drug in and out of the micelle is also greatly influenced <strong>by</strong> the<br />

properties of the micelle core. In general, increasing the molecular weight of the core-forming<br />

block will increase the size of the core provided that the hydrophilic block length is held constant.<br />

This increase in the molecular weight of the hydrophobic block and the size of the core has been<br />

found to decrease the overall rate of release rate of the entrapped agent. 63,68–71<br />

The nature of the core, specifically the hydrophobicity and/or polarity of the core, influences the<br />

release rate of the drug as it determines the permeability of the core to aqueous media. Micelles with<br />

a highly hydrophobic core will likely have less aqueous fluid in the core in comparison to micelles<br />

that have a more hydrophilic or polar core. Therefore, micelles prepared from more hydrophilic<br />

cores will likely have an accelerated rate of drug release. The physical state of the core will also<br />

affect the diffusion of drug from the micelles. Polymers that have a high T g or bulky groups present<br />

on their backbone are limited in terms of their ability to reorient. As a result, these polymers will<br />

form micelle cores with a high microscopic viscosity, resulting in a low diffusion coefficient for the<br />

drug that is incorporated because of limited movement. 72 This is in contrast to polymers that are<br />

more rubbery or gel-like that will tend to form cores that have a low microscopic viscosity and a<br />

higher rate of diffusion for the incorporated drugs.<br />

Finally, the degree of polymer–drug compatibility greatly influences the rate of drug release<br />

from the micelles. Drugs that are highly miscible with the core-forming block have a high degree of<br />

polymer–drug compatibility. It is expected that these drugs would be molecularly dissolved in the<br />

core and not exhibit any crystallinity such that the release would be relatively fast. However, it is<br />

necessary to consider that the actual effect of a high degree of miscibility on drug release depends<br />

on the strength and extent of the core–drug interactions. 36,45 An increase in the extent of interaction<br />

between a drug and the polymer will result in a decrease in the diffusion coefficient of the drug<br />

(slower release). 73 The ability to provide sustained release of the drug from micelle systems is<br />

particularly beneficial for their use in therapeutic applications.<br />

17.3 PHYSICAL CHARACTERISTICS OF POLYMERIC MICELLES FOR<br />

CLINICAL APPLICATIONS IN CANCER THERAPY<br />

In the past two decades, many groups have reported on the development and optimization of<br />

polymeric micelles for the delivery of anti-cancer drugs. In particular, much effort has focused<br />

on the formulation of doxorubicin and paclitaxel in micelles because of their dose-limiting toxicities<br />

in currently approved formulations and high potency against a wide variety of cancers. To<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 327<br />

(a)<br />

O<br />

OCH 3 O<br />

date, several doxorubicin- (i.e., NK911, SP1049C) and paclitaxel-loaded (i.e., Genexol-PM,<br />

NK105) micelle formulations have entered clinical trial development.<br />

17.3.1 DOXORUBICIN<br />

Doxorubicin (Adriamycin w ), a member of the anthracycline antibiotics family, is currently<br />

employed for treatment of many types of cancer because of its broad anti-tumor spectrum. It is<br />

approved, either alone or in combination with other chemotherapeutic drugs, for treatment of<br />

breast, ovarian, bladder, lung, thyroid, and gastric cancer as well as malignant lymphoma, acute<br />

forms of leukemia, and sarcomas. 74 The cytotoxic effect of doxorubicin is said to result from a<br />

combination of cell damage mechanisms, some that have not been conclusively identified. One of<br />

the effects of doxorubicin that is known to contribute to its anti-tumor activity is the inhibition of<br />

cellular replication and DNA transcription as a result of an irreversible change in the DNA tertiary<br />

structure. 75 This non-cell type specific cytotoxic effect of doxorubicin is mainly attributed to its<br />

ability to intercalate between DNA base pairs with its planar structure (Figure 17.3a). Indeed, the<br />

planar moiety of doxorubicin has been shown to improve the stability of micelles when loaded in<br />

the micelle core. 76 Doxorubicin has been successfully loaded into micelles formed from PEG-bpoly(3-caprolactone)<br />

(PEG-b-PCL), 41 PEG-b-poly(D,L-lactide-co-glycolide) (PEG-b-PLGA), 77,78<br />

PEG-b-poly(propylene oxide)-b-PEG (PEG-b-PPO-b-PEG; Pluronic w ), 26,79–81 and PEG-b-poly<br />

(aspartic acid) (PEG-b-PAsp). 16,32,82,83 Currently, two doxorubicin-loaded micelle formulations,<br />

NK911 (i.e., PEG-b-PAsp) and SP1049C (i.e., Pluronic w ) are being investigated in clinical trials.<br />

17.3.1.1 NK911<br />

H 3C<br />

OH<br />

OH<br />

O<br />

O<br />

OH<br />

O<br />

CH2OH OH<br />

NH 2<br />

NK911 is a doxorubicin-loaded PEG-b-PAsp copolymer micelle formulation developed <strong>by</strong><br />

Kataoka and co-workers. Doxorubicin is both physically entrapped in the micelle core and chemically<br />

conjugated to the aspartic acid side chains of the core-forming block via amide linkages.<br />

NK911 is currently undergoing phase II clinical trial evaluation for efficacy and toxicity profiles.<br />

The formulation entered phase I in 2001, and the results from this trial revealed that NK911 exhibits<br />

a longer circulation half-life, a larger area under the curve (AUC), and reduced toxicities in<br />

comparison to the conventional formulation of doxorubicin. 82 The prolonged circulation lifetime<br />

of NK911 is attributed to its composition and the monomodal size distribution of the micelles with a<br />

mean diameter of approximately 50 nm. 16,82<br />

O<br />

O OH<br />

FIGURE 17.3 Structures of anti-cancer drugs, (a) doxorubicin and (b) paclitaxel, that have been formulated in<br />

polymeric micelle formulations for clinical trial development.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

(b)<br />

O<br />

NH<br />

OH<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

O<br />

H<br />

O<br />

O<br />

O


328<br />

The amphiphilic or polar nature of doxorubicin (log P value of 0.52) facilitates the partitioning<br />

of this drug from the micelle core into the external medium. 84 Therefore, chemical conjugation of<br />

doxorubicin to the core-forming block can enhance drug retention and retard drug release from the<br />

micelles. The initial development of the PEG-b-Asp micelles for drug delivery consisted of only<br />

conjugated doxorubicin to the PAsp core. Specifically, PEG-b-PAsp copolymers were synthesized<br />

from PEG-b-PBLA with the subsequent addition of doxorubicin to the carboxyl groups on the side<br />

chains of the core-forming block. As a result of the conjugation, a 37 mol% substitution of doxorubicin<br />

was achieved that is equivalent to 20 mg/mL doxorubicin. 85 The doxorubicin residues of the<br />

conjugates stabilize the micelle core through pi–pi stacking and hydrophobic interactions,<br />

providing the micelles with excellent kinetic stability in vivo. 76<br />

Subsequent studies revealed that the chemically conjugated doxorubicin exerts negligible antitumor<br />

activity in vitro and in vivo as a result of the lack of a cleavable linkage between the drug and<br />

the copolymer. Therefore, the next generation of the PEG-b-PAsp micelles was prepared to include<br />

physically encapsulated doxorubicin within the micelle core as well as the chemically conjugated<br />

drug. The amount of doxorubicin that could be physically entrapped within the core was found to be<br />

proportional to the amount of conjugated doxorubicin present. The enhanced compatibility between<br />

the physically entrapped drug and the micelle core is attributed to the presence of the chemically<br />

conjugated drug. In vivo studies confirmed that micelles with the highest amount of physically<br />

entrapped doxorubicin had the most potent anti-cancer activity. 32<br />

The formation of a dimerized form of the physically entrapped doxorubicin was detected within<br />

the micelles. Despite the fact that the dimer was found to have insignificant anti-tumor activity<br />

in vitro and in vivo, it plays a supplementary role to stabilize the entrapped doxorubicin. The<br />

addition of dimers to the micelles as part of the physically entrapped components can further<br />

prolong the circulation lifetime of the micelles in vivo. 86 However, the addition of doxorubicin<br />

dimers was also found to decrease the aqueous solubility of the drug-loaded micelles that compromised<br />

the stability of the micelles following lyophilization and long-term storage. Therefore, the<br />

fully optimized NK911 formulation was prepared such that it only contained doxorubicin<br />

monomer. 82<br />

17.3.1.2 SP1049C<br />

Nanotechnology for Cancer Therapy<br />

SP1049C is a doxorubicin-loaded Pluronic w copolymer micelle formulation developed <strong>by</strong> Kabanov’s<br />

group and Supratek Inc. (Montreal, Canada). In this formulation, doxorubicin is physically<br />

encapsulated in micelles assembled from two different types of Pluronic w copolymers, L61 and<br />

F127 in a 1:8 w/w ratio. 26 In October 2005, SP1049C was granted orphan drug status <strong>by</strong> the FDA<br />

for the treatment of oesophageal carcinoma based on results from the phase I study; meanwhile, its<br />

phase II clinical trial results are currently under final review. In contrast to NK911, the phase I<br />

clinical trial of SP1049C deliberately recruited patients with cancers that were refractory following<br />

conventional treatment and patients with previous anthracycline treatment. 81 This is due to the<br />

established chemosensitizing effect of SP1049C against multi-drug resistant (MDR) cancers, an<br />

effect that is attributed to its Pluronic w copolymer composition.<br />

The Pluronic w copolymers are the class of copolymers that have been most widely evaluated<br />

and employed in pharmaceutical applications. Studies have indicated that Pluronic w copolymers<br />

having an hydrophile-lipophile balance (HLB) of less than 19 and a moderate molecular weight are<br />

most effective in overcoming MDR. 34 In a separate study, Pluronic w L61 that has an HLB value of<br />

7 was found to be the most effective chemosensitizer among other Pluronic w copolymers. 26,34<br />

Doxorubicin loaded in Pluronic w L61 micelles was found to be 2–3 orders of magnitude more<br />

toxic than conventional doxorubicin when evaluated in a range of drug-resistant cancer cells (e.g.,<br />

MCF-7/ADR breast carcinoma, SKVLB ovarian carcinoma, KBV oral epithelial carcinoma, LoVo/<br />

Dox colon adenocarcinoma, P388-Dox murine leukemia cells, SP2/0Dox myeloma as well as<br />

others). 26 Furthermore, mechanistic studies have shown that it is the unimers rather than the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 329<br />

micelles that are responsible for the reversal of MDR. 26,34 More details on the mechanisms and<br />

effects of Pluronic w as chemosensitizers are given in Section 17.4.<br />

Despite the effectiveness of micelles formed from only Pluronic w L61, the hydrophobicity of<br />

the copolymer results in unfavorable stability issues. Phase separation and micelle aggregation<br />

leading to microembolic effects have been observed in preliminary in vivo toxicity studies of the<br />

L61 formulation. Kabanov and co-workers solved this problem <strong>by</strong> introducing a higher molecular<br />

weight and more hydrophilic Pluronic w copolymer (F127) into the L61 micelle formulation. By<br />

varying the ratio of the L61 to the F127 copolymer, it was found that at 0.25 w/w% L61 (i.e.,<br />

effective dosing concentration of L61) and 2.0 w/w% F127, the Pluronic w micelles maintained their<br />

effective diameter of 22.4 nm with no aggregation. 26 In addition, the IC 50 of doxorubicin loaded in<br />

the L61/F127 micelles was similar to that of the L61 formulation. Therefore, the optimized<br />

SP1049C formulation utilized for the clinical studies contains 2.0 mg/mL doxorubicin, 2.5 mg/<br />

mL L61, and 20 mg/mL F127 in 0.9% sodium chloride. 81<br />

17.3.2 PACLITAXEL<br />

Similar to doxorubicin, paclitaxel is a chemotherapeutic drug that has been loaded into block<br />

copolymer micelles for investigation in clinical trials. Paclitaxel is an anti-neoplastic drug that<br />

was first extracted from the bark of Pacific yew trees (Figure 17.3b). The cytotoxic effect of<br />

paclitaxel is mainly attributed to its ability to promote the assembly as well as prevent the depolymerization<br />

of microtubules. 87 Stabilization of the microtubule networks inhibits normal<br />

interphase and mitotic functions that, in turn, impedes cell division. Paclitaxel is approved for<br />

treatment of ovarian, breast, and non-small cell lung cancers as well as AIDS-related Kaposi’s<br />

sarcoma. 87 Paclitaxel is highly hydrophobic with a log P value of 3.96 and low aqueous solubility<br />

of 1 mg/mL. 88,89 The commercialized formulation of paclitaxel, known as Taxol w , contains the<br />

drug dissolved in a mixture of Cremophor EL and ethanol that increases its solubility <strong>by</strong> more than<br />

6000-fold. However, administration of Cremophor EL leads to severe dose-limiting toxicities such<br />

as hypersensitivity reactions and neuropathy. 90 As a result, much effort has gone into the development<br />

of a novel formulation of paclitaxel using biocompatible and biodegradable polymeric<br />

micelles. Paclitaxel has been successfully loaded into PEG-b-poly(D,L-lactide) (PEG-b-<br />

PDLLA), 35,37,90,91 poly(N-vinyl-pyrrolidone)-b-PDLLA (PVP-b-PDLLA), 28,92 PEG-b-PCL, 93–95<br />

and PEG-b-poly(d-valerolactone) (PEG-b-PVL). 95 Genexol-PM (i.e., PEG-b-PDLLA) and<br />

NK105 (i.e., PEG-b-PAsp) represent two paclitaxel-loaded polymeric micelle formulations that<br />

have entered clinical trial development.<br />

17.3.2.1 Genexol-PM<br />

Genexol-PM, developed <strong>by</strong> Samyang Co. (Korea), is a paclitaxel formulation based on PEG-b-<br />

PDLLA micelles. In 2004, results from a phase II clinical trial that evaluated Genexol-PM and<br />

cisplatin as treatment for patients with advanced gastric cancer was reported. 96 In the first clinical<br />

trial, the AUC and Cmax of Genexol-PM was found to be lower than that of Taxol w ; however, the<br />

maximum tolerated dose (MTD) was increased from 135 to 390 mg/m 2 . 90 Preclinical studies also<br />

revealed similar pharmacokinetic profiles for Genexol-PM and Taxol w in the B16 melanoma<br />

murine model. There are two explanations for the above observations. First, the liquid-like core<br />

of PDLLA acts as a solubilization site for paclitaxel as opposed to an actual drug carrier, 35 and<br />

second, the decrease in AUC and Cmax occurs as a result of the shift in the drug accumulation to the<br />

tumor tissues that was evidenced in preclinical biodistribution studies. 97 Formulation development<br />

of paclitaxel-loaded PEG-b-PDLLA micelles was mostly reported <strong>by</strong> Burt and co-workers. 35,37,91<br />

Genexol-PM is prepared <strong>by</strong> the evaporation method that involves co-dissolving paclitaxel and<br />

PEG-b-PDLLA in acetonitrile with subsequent evaporation of the solvent to form a copolymer–<br />

drug matrix. The matrix is then heated to 608C with the addition of water or buffer for self-assembly<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


330<br />

of paclitaxel-loaded micelles. 97 The evaporation method is much more suitable than the direct<br />

dissolution method because paclitaxel and PEG-b-PDLLA are relatively insoluble in aqueous<br />

media. Indeed, the method employed for preparation of the paclitaxel-loaded PEG-b-PDLLA<br />

micelles can influence the extent of drug loading and physical stability of the formulation. The<br />

choice of solvent strongly influences the extent to which paclitaxel can be solubilized in PEG-b-<br />

PDLLA micelles as reported <strong>by</strong> Zhang et al. 37 Among many of the solvents tested, only acetonitrile<br />

resulted in a clear micelle solution with no evidence of drug precipitation. It was suggested that<br />

acetonitrile enhances the miscibility between paclitaxel and the PDLLA block that, in turn, reduces<br />

the possibility of phase separation during the solvent evaporation process. 37<br />

Burt and co-workers also studied the use of PEG-b-poly(D,L-lactide-co-caprolactone) and PEGb-poly(glycolide-co-caprolactone)<br />

copolymers for encapsulation of paclitaxel. The addition of<br />

hydrophobic caprolactone to PDLLA could potentially increase paclitaxel loading through the<br />

increased hydrophobicity of the core-forming block. However, these micelles were found to<br />

have poor physical stability and resulted in the precipitation of paclitaxel. 35 The PEG-b-PDLLA<br />

micelles can solubilize as much as 5% (w/v) paclitaxel and remain stable for up to 24 h. This is<br />

mainly attributed to the absence of crystalline paclitaxel in the paclitaxel/PEG-b-PDLLA matrix as<br />

revealed <strong>by</strong> x-ray diffraction studies. In fact, paclitaxel is molecularly dissolved in the PDLLA core<br />

that prevents drug precipitation and results in excellent solubilization of the drug. 37,91<br />

The amorphous state of the paclitaxel–PDLLA core can be problematic for in vivo applications.<br />

Studies have indicated that the presence of crystalline drug in the core can retard drug release. 69,98<br />

Therefore, the amorphous core of the paclitaxel-loaded PEG-b-PDLLA micelles can lead to a much<br />

faster drug release profile. As demonstrated <strong>by</strong> Burt et al. following intravenous administration<br />

paclitaxel rapidly dissociates from the micellar components in the circulation. 35 Therefore, it is<br />

likely that Genexol-PM acts as a solubilizer for paclitaxel rather than as a true drug carrier.<br />

17.3.2.2 NK105<br />

Nanotechnology for Cancer Therapy<br />

In 2005, Kataoka and co-workers, in collaboration with NanoCarrier Co., Ltd. and Nippon Kayaku<br />

Co., Ltd., first reported on NK105, a paclitaxel-incorporated micelle formulation. 99 The phase I<br />

clinical trial of NK105 first began in April 2004. Similar to NK911, the building block of NK105 is<br />

PEG-b-PAsp copolymers but with a slight modification to the aspartate blocks such that the<br />

carboxyl residues are converted into 4-phenyl-1-butanolate <strong>by</strong> an esterification reaction. The<br />

modification increases the hydrophobicity of the core-forming block that results in enhanced<br />

entrapment and retention of paclitaxel in the micelles. The resulting drug-loaded micelles<br />

contain 23% (w/w) paclitaxel with a weight–average diameter of approximately 85 nm.<br />

With the improved micelle core stability, NK105 is predicted to act as a true drug carrier rather<br />

than a solubilizer. As indicated <strong>by</strong> the in vivo preclinical studies, NK105 is extremely stable in vivo<br />

with a 90-fold increase in the AUC and a 4–6-fold increase in the terminal phase half-life. 99 In<br />

addition, the tumor AUC was found to be 25-times higher than that of Taxol w . The superior in vivo<br />

anti-tumor activity of NK105 in a human colorectal cancer xenograft model (HT-29) was a result of<br />

the enhanced drug accumulation in the tumor, allowing for more drug-loaded micelle–cell<br />

interactions. These results highlight the importance of polymer–drug compatibility and core stability<br />

in the design of effective micellar delivery systems. Therefore, the recent studies that focus on<br />

customizing copolymer compositions to accommodate the distinct properties of different drugs<br />

such as camptothecin, 38,100 KRN 5500, 60 and cisplatin 101,102 may lead to the development of true<br />

polymeric micelle carriers for delivery to tumor tissues.<br />

17.4 INTERACTIONS OF POLYMERIC MICELLES WITH CANCER CELLS<br />

Once the drug-loaded micelles have reached the tumor site, there are several mechanisms that<br />

have been proposed for the delivery of micelle-incorporated drugs into the cancer cells as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 331<br />

(a)<br />

(b)<br />

illustrated in Figure 17.4. 103–106 In general, drugs can enter the cells via diffusion as free<br />

molecules following release from the micelles (Figure 17.4a and b) or be internalized as drug<br />

molecules entrapped within the micelles (Figure 17.4c–f). In addition, certain block copolymers<br />

have also been found to act as chemosensitizers or modulators of drug efflux pumps. 26,34,107–111<br />

17.4.1 DRUG DIFFUSION<br />

Lysosome<br />

Nucleus<br />

Lysosome<br />

Endosome<br />

FIGURE 17.4 Proposed mechanisms for the delivery of micelle-formulated drugs into cancer cells. (a) Drug is<br />

released from the micelles and diffused into the cell as free drug. (b) The drug-loaded micelles diassemble in<br />

the extracellular space and release the drug; the released drug then diffuses across the cell membrane into the<br />

intracellular space. (c) Intact drug-loaded micelles are internalized via adsorptive pinocytosis after adsorption<br />

onto the cell surface or (d) <strong>by</strong> fluid-phase endocytosis. The internalized drug-loaded micelles eventually reach<br />

the lysosome where the drug can be released (e) <strong>by</strong> diffusion or (f) following disassembly or degradation of the<br />

micelles.<br />

A free drug can diffuse across the cell membrane following disassembly of the micelles or release<br />

of the drug from intact micelles. The extent and rate of diffusion of the free molecules across the<br />

cell membrane will depend largely on the physico-chemical properties of the drug (water solubility,<br />

charge, molecular weight, and partition coefficient). Small hydrophobic molecules can easily<br />

permeate the membrane and enter the intracellular space. Studies have shown that the rate of<br />

cell uptake of a free drug (non-micelle-incorporated) is faster than that of drug incorporated in<br />

micelles. 103–105 In this way, the release profile of the drug from the micelles can have a significant<br />

influence on the rate of cellular uptake of the drug. As the drug is released from the micelles, there is<br />

an increase in the gradient between the concentrations of free drug on the exterior and interior of the<br />

cell that, in turn, drives diffusion of the drug into the intracellular compartment. A micelle system<br />

with a slow drug release profile will only have a limited pool of free drug available for diffusion into<br />

the cell. For example, Maysinger et al. demonstrated that for PEG-b-PCL micelles loaded with<br />

benzo[a]pyrene that has an apparent partition coefficient (Kv) of 690, the rate of cell uptake of the<br />

micelle formulated probe was similar to that of the rate of uptake of free benzo[a]pyrene. In<br />

contrast, for micelles loaded with DiI that has a Kv of 5800, a much lower degree of cell uptake<br />

of the micelle formulated probe was observed when compared to free DiI. 45,103 Therefore, at the<br />

site of interest, an increase in the rate of drug release from the micelles can result in a greater extent<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

(e)<br />

(c)<br />

(f)<br />

(d)


332<br />

of intracellular drug accumulation if the drug can easily diffuse across the cell membrane. Micelle<br />

systems that respond to external stimuli as a means to promote drug release at the tumor have also<br />

been designed and are discussed further in Section 17.6. Another route for the entry of micelleformulated<br />

drug into cells includes the internalization or endocytosis of the drug-loaded<br />

micelles. 103,104,106<br />

17.4.2 CELLULAR INTERNALIZATION<br />

Macromolecules and nanosized particles including polymer–drug conjugates, liposomes, and<br />

micelles are unable to diffuse through the cell membrane into the intracellular compartment. In<br />

order to enter the cell, these drug delivery vehicles require an internalization mechanism such as<br />

phagocytosis or pinocytosis. Only certain cell types called phagocytic cells (macrophages, neutrophils)<br />

can undergo phagocytosis. Particulate materials with diameters of less than 1 mm can be<br />

internalized into these cells in membrane-bound vesicles called phagosomes. 112 Conversely, pinocytosis<br />

is known to be non-cell type specific. This mechanism occurs <strong>by</strong> invagination of the cell<br />

membrane that is similar to the mechanism in phagocytosis. However, the pinocytic vesicles that<br />

pinch off from the cell membrane are generally smaller than phagosomes and have diameters<br />

ranging from 100 to 200 nm. 112 The pinocytosis mechanism can be further categorized into<br />

fluid-phase and adsorptive.<br />

Fluid-phase pinocytosis involves the engulfment of soluble materials and extracellular fluid<br />

during continual cell membrane ruffling. This is in contrast to adsorptive pinocytosis that involves<br />

the association of particulate materials with cell membrane components prior to internalization. 113<br />

(e)<br />

(a)<br />

k d<br />

k a<br />

(d)<br />

Early endosome<br />

(b)<br />

(f)<br />

Late endosome<br />

Nanotechnology for Cancer Therapy<br />

(c)<br />

Nucleus<br />

FIGURE 17.5 Internalization of ligand-conjugated micelles via receptor-mediated endocytosis. (a) The<br />

ligand-coupled micelles approach the cancer cells and (b) bind to the cell surface receptors depending on<br />

the association constant (ka) and the dissociation constant (kd) of the receptor–ligand complexes. (c) The<br />

micelles are then rapidly internalized into the cells via receptor-mediated endocytosis. (d) The internalized<br />

receptor and the ligand-conjugated micelles then undergo endosomal sorting in the early endosomes. In most<br />

cases, (e) the receptor is recycled to the cell surface, and (f) the micelles enter the late endosome that eventually<br />

fuses with the lysosome.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 333<br />

For example, micelles that are conjugated with specific ligands can bind to the corresponding<br />

receptors that are expressed on the cell surface; as a result, the micelles can be internalized via<br />

receptor-mediated endocytosis. 78,114,115 More details on receptor-mediated endocytosis are<br />

provided in Section 17.6 and Figure 17.5. In most cases, micelles are internalized into the cells<br />

through fluid-phase endocytosis. 77,116,117<br />

Fluid-phase endocytosis is expected to result in a lower degree of cell uptake in comparison to<br />

adsorptive pinocytosis or diffusion of free hydrophobic agents. Differences in the cell uptake of<br />

micelle-incorporated agents and non-micelle-incorporated agents have been demonstrated in<br />

various studies. For example, using a hydrophobic fluorescent probe as a tracker for micelles,<br />

the rate and amount of intracellular fluorescence was found to be much lower than incubation<br />

with free probes. 104 Thishasalsobeenconfirmed<strong>by</strong>otherstudies that relied on confocal<br />

microscopy as a qualitative tool. 103,105 Rapoport et al. also demonstrated that the incorporation<br />

of doxorubicin and ruboxyl into Pluronic w micelles resulted in a decrease in the extent of cell<br />

uptake of the agents in ovarian cancer cells. 118 In a drug accumulation study at copolymer concentrations<br />

above the CMC of the Pluronic w copolymers, the accumulation of the fluorescent probe<br />

R123 was substantially decreased in both non-multi- and multi-drug resistant cancer cells. It was<br />

suggested that the R123-loaded micelles were transported into the cells via a vesicular route and<br />

followed accumulation kinetics different from that of the non-micelle incorporated R123. 34<br />

Endocytosis is known to be an energy-, time-, and pH-dependent process. 113 Many pharmacological<br />

manipulations have been employed in order to demonstrate that micelles are internalized<br />

via an endocytotic pathway. For example, a low pH and incubation at 48C resulted in the reduction<br />

of the internalization of micelles. 104 Also, pre-treatment of cells with metabolic inhibitors such as<br />

sodium azide and 2-deoxyglucose (DOG) that deplete the energy source for membrane ruffling<br />

resulted in a decrease in the extent of uptake of radiolabeled-drug loaded-PEG-b-PCL micelles <strong>by</strong><br />

at least 80%. Another study conducted <strong>by</strong> Luo et al. reported similar findings, yet in this case, the<br />

micelles were tracked using fluorescently labeled-copolymer. 106 Furthermore, the pre-treatment of<br />

cells with Brefeldin A, known to alter the morphology of endosomes, was also found to reduce the<br />

extent of uptake of micelles. This suggests that an endosomal compartment may be involved in the<br />

internalization of the micelles that further supports the claim that micelle internalization does<br />

proceed <strong>by</strong> an endocytotic mechanism. 104<br />

In order to confirm that micelles have been internalized via the endocytotic pathway, different<br />

methods have been developed for the detection of micelles in the relevant organelles. For example,<br />

a pH-sensitive fluorescent probe has been used to confirm the presence of the micelles in acidic<br />

organelles using flow cytometry. 118,119 The subcellular localization of micelles can also be visualized<br />

using confocal microscopy <strong>by</strong> fluorescently labeling the micelles and counterstaining the<br />

intracellular organelles. 41,105,115 Hydrophobic fluorescent probes physically entrapped in micelles<br />

can also be used to track the micelles. The use of these probes to follow the pathway or fate of the<br />

micelles requires virtually no or minimal release of the probe during the time frame of the experiments.<br />

However, cell culture media often contain protein that can act to accelerate drug release. 120<br />

Therefore, the accuracy associated with using physically entrapped probes to track micelles is<br />

questionable because of the different subcellular localization patterns of free and micelle-incorporated<br />

drugs. 41,105 Ideally, fluorescent probes that are chemically conjugated to the copolymers<br />

provide the most reliable tool for following the micelles in vitro. Although copolymers also<br />

exist in unimeric form, the concentration of unimers in a micelle solution does not exceed the<br />

CMC of the copolymer and represents only a minor population. 106 Therefore, it is reasonable to<br />

assume that the fluorescence detected corresponds to the micelle population provided that the<br />

conjugation of the probe to the copolymers is stable.<br />

Internalized micelles have been shown to localize mainly in cytoplasmic compartments, in<br />

particular, the accumulation of micelles in acidic organelles (endosomes or lysosomes) has been<br />

observed. 105,118,119 Rapoport et al. have reported on the intracellular uptake and trafficking of<br />

Pluronic w micelles. Their results suggested that after the micelles were internalized in acidic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


334<br />

organelles, the non-ionic Pluronic w copolymers induced permeabilization of the membranes,<br />

resulting in the release of the encapsulated drugs into the cytosol or into the nucleus via diffusion<br />

of the drug. 118 The nuclear pore complexes on the nucleus allow for passive diffusion of small<br />

molecules that have diameters of less than 9 nm, whereas macromolecules as large as 39 nm can be<br />

actively transported into the nucleus. 121 Theoretically, micelles that are greater than 9 nm cannot be<br />

translocated into the nucleus unless nuclear targeting signals are present on the particles. For<br />

example, in cancer cells that have been incubated with tetramethylrhodamine-5-carbonyl azide<br />

(TMRCA)-conjugated PEG-b-PCL micelles for 24 h, only cytoplasmic localization of the micelles<br />

was observed, and the nuclear compartment was devoid of micelles. 105 Similar results were<br />

obtained <strong>by</strong> Shuai et al. with doxorubicin-loaded PEG-b-PCL micelles that have a delayed<br />

release profile for doxorubicin from the micelles. The authors demonstrated that following incubation<br />

of cells with micelle formulated doxorubicin for 2 and 24 h, the drug was only localized in<br />

cytoplasmic compartments, whereas incubation of free drug with cells resulted in nuclear accumulation.<br />

41 To date, the cellular internalization pathway and fate of micelles and micelle-encapsulated<br />

drugs as a function of time have not been clearly elucidated. However, the in vitro pathway and fate<br />

of micelles can have a significant impact on the cytotoxic effects of the formulated drug. The<br />

various anti-cancer agents that have been traditionally formulated in micelles have different sites of<br />

action within the cell. For example, paclitaxel and doxorubicin are required to bind to microtubules<br />

in the cytoplasm and DNA in the nucleus, respectively, in order to exert their cytotoxic effects.<br />

Therefore, if micelles could be engineered or designed to achieve a specific subcellular localization<br />

that matched the required localization of the drug, the cytotoxicity of the drug could be enhanced.<br />

17.4.3 IN VITRO CYTOTOXICITY<br />

Nanotechnology for Cancer Therapy<br />

The subcellular localization and the release profile of drug-loaded micelles can impact the cytotoxic<br />

effects of the drugs. In most cases, free drugs have been found to have lower IC 50 values than drugs<br />

formulated in micelles. Such an increase in the IC 50 of the micelle-formulated drugs is likely<br />

because of the delayed release of drugs from the micelle core. 41,95,122<br />

Shuai et al. reported delayed cell death in MCF-7 breast cancer cells when using a doxorubicin<br />

formulated PEG-b-PCL micelle system in comparison to free doxorubicin. 41 The confocal<br />

microscopy images showed that the micelle-incorporated doxorubicin resides in the cytoplasm,<br />

and free doxorubicin accumulates quickly in the cell nucleus. The delayed cell death could be due<br />

to the difference in the subcellular localization of free and micellized drug because doxorubicin can<br />

only exert its cytotoxic effect after reaching the nucleus. 41 Interestingly, as reported <strong>by</strong> Yoo and<br />

Park, doxorubicin-conjugated PEG-b-PLGA micelles, led to a 10-fold increase in the cytotoxicity<br />

of this drug. However, in this case, doxorubicin was found to localize in both the cytoplasm and the<br />

nucleus when delivered <strong>by</strong> micelles. 77 The discrepancy between the intracellular localization of<br />

doxorubicin formulated in the two distinct systems may be explained <strong>by</strong> their different drug release<br />

profiles. The PCL core is more hydrophobic than the PLGA core and can, therefore, provide better<br />

drug retention. Although doxorubicin was conjugated to the PLGA system that should result in a<br />

much slower release, the hydrolytic linkage is susceptible to cleavage and results in a faster release<br />

of the drug when compared to the PCL system. The cytotoxicity observed in the PEG-b-PLGA<br />

micelle system can be attributed to the doxorubicin that has been cleaved from the copolymer and<br />

released from the micelles because the conjugated drug was found to be non-cytotoxic. 16,32<br />

Furthermore, using a thermosensitive micelle system, Liu et al. obtained accumulation of<br />

doxorubicin in the cytoplasm and the nucleus following a temperature increase, and no nuclear<br />

accumulation of the drug was observed prior to the increase in temperature. 122 The delayed cytotoxicity<br />

of micelle-formulated drugs may seem to be a limitation in terms of the utilization of this<br />

technology as cancer therapy. However, the retention of drug in the micelles in vivo is required in<br />

order to ensure successful delivery of the drug to the tumor site. The relationship between drug<br />

retention in micelles and in vivo efficacy is discussed in Section 17.5.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 335<br />

Copolymer micelles have also been found to amplify the effect of chemotherapeutic drugs in<br />

cancer cells. In the case of multi-drug-resistant cancer cells, certain copolymers can also act as<br />

chemosensitizers and increase the intracellular accumulation of the drug. The Pluronic w copolymers<br />

are the family of materials that have been investigated most extensively in terms of their role<br />

as chemosensitizers or modulators of drug efflux.<br />

17.4.4 INFLUENCE OR EFFECT ON MULTI-DRUG RESISTANCE<br />

Multi-drug resistance (MDR) is one of the most significant obstacles in chemotherapy and is often<br />

found in refractory cancers. MDR may be a result of the undesired efflux of a drug from the<br />

intracellular compartment. Transporter proteins that are overexpressed on the cell membrane of<br />

certain cancer cells are responsible for drug efflux. The drug efflux protein that has been studied<br />

most extensively is the ATP-dependent P-glycoprotein (P-gp). 26,123 This drug transporter is also<br />

expressed in healthy tissues such as the epithelial cells of the intestine and the endothelial cells in<br />

the blood–brain barrier. The normal physiological function of P-gp is to regulate molecular transport<br />

and provide protection or detoxification of the cell. 124 The normal physiological levels of drug<br />

efflux protein also function to limit the bioavailability of drugs delivered via the oral route as well as<br />

to limit drug delivery to the brain.<br />

Kabanov’s group has made significant contributions in the development of Pluronic w micelles<br />

as a tool for enhancing drug delivery both across the blood–brain barrier and into MDR<br />

tumors. 7,26,34,108,110,124–127 Their work in this area originated from the finding that doxorubicin<br />

formulated in Pluronic w micelles was more cytotoxic than free doxorubicin in drug-resistant cells.<br />

Pluronic w can increase the cytotoxicity of anthracyclines in drug-resistant cancer cells <strong>by</strong> 2–3<br />

orders of magnitude. For instance, the IC50 of daunorubicin in the presence of Pluronic w copolymers<br />

was decreased <strong>by</strong> 700-fold in resistant ovarian cancer cells (SKVLB). 125 Additional results on<br />

the cytotoxicity of anti-cancer drugs formulated in Pluronic w micelles in resistant cancer cells are<br />

summarized elsewhere. 108,125 Further mechanistic studies have demonstrated that the ability of<br />

Pluronic w copolymers to inhibit drug efflux proteins depends on their composition.<br />

Specific compositions of Pluronic w copolymers were evaluated in a screening study in order to<br />

demonstrate that the inhibitory effects of the copolymers on drug efflux were dependent on the<br />

length of the PEG and PPO segments. It was found that hydrophobic Pluronic w copolymers with<br />

values of HLB!19 and with an intermediate length of the PPO segments such as L61 (PEG2–<br />

PPO30–PEG2) and P85 (PEG26–PPO40–PEG26) are the most effective at inhibiting P-gp. 34<br />

SP1049C is a Pluronic w -based micelle formulation that has been selected for further preclinical<br />

development and has entered phase II clinical trial for the treatment of relapsed cancers. L61 was<br />

chosen as the base component for the SP1049C formulation with the addition of F172 (another high<br />

molecular weight hydrophilic Pluronic w ) copolymers to stabilize the formulation as discussed in<br />

Section 17.3. 26<br />

In many of the studies, the P85 copolymer was fundamental in the investigation of the<br />

mechanisms of Pluronic w copolymers for overcoming drug resistance in cancer. 23,108,109,124 To<br />

date, several mechanisms have been hypothesized to account for the role that Pluronic w copolymers<br />

play as drug efflux modulators, and these have been reviewed in detail elsewhere. 26,123,125 It has<br />

been confirmed that the unimers, as opposed to the copolymer micelles, are responsible for inhibiting<br />

drug efflux. For example, accumulation and cytotoxicity of doxorubicin were demonstrated<br />

to increase with an increase in the Pluronic w concentration up to the CMC of the copolymer in drugresistant<br />

human breast cancer cells (MCF-7/ADR) as well as head and neck cancer (KBv) cells.<br />

However, after reaching the CMC of the copolymers (the concentration when micelles are formed),<br />

both the intracellular doxorubicin content and the cytotoxicity were found to level off and/or<br />

decrease. 34 Similarly, the extent of intracellular accumulation of fluorescein, a multi-drug resistance-associated<br />

protein (MRP) substrate, in an MRP-expressing human pancreatic adenocarcinoma<br />

cell line (Panc-1) was found to be dependent on the concentration of Pluronic w present. 127<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


336<br />

Nanotechnology for Cancer Therapy<br />

The ability of Pluronic w unimers to inhibit drug efflux transporters is mainly attributed to ATPdepletion.<br />

26,109 Use of certain metabolic inhibitors such as sodium azide can inhibit P-gp and result<br />

in increased drug accumulation <strong>by</strong> depleting the energy source for efflux across the cell membrane.<br />

Similar effects are obtained with the addition of Pluronic w copolymers to the cells as they result in<br />

an abolishment of the energy supply within the MDR cells. 34,109,125,127 For example, P85 and P105<br />

were found to reduce the activity of the electron transport chain reactions in the mitochondria in<br />

HL-60 cells, decreasing the degree of oxidative metabolism, indicating that Pluronic w copolymers<br />

can interrupt the production of ATP. 128 In addition, cell exposure to P85 has been found to induce<br />

reversible transient ATP depletion in Jurkat lymphoma T-cells. 129 These effects are much more<br />

profound in MDR cells because the drug efflux mechanisms are ATP-driven. 109 It was believed that<br />

MDR cells consume energy at a much higher rate than non-MDR cells; therefore, their activity is<br />

more susceptible to inhibition via ATP-depletion. As a result, the Pluronic w copolymers were<br />

capable of chemosensitizing the MDR cells to a greater extent than non-MDR cells. All of the<br />

above suggest that Pluronic w -induced ATP depletion is the major mechanism that is responsible for<br />

inhibiting drug efflux in MDR cells.<br />

Another proposed mechanism of drug efflux inhibition is membrane fluidization caused <strong>by</strong><br />

Pluronic w copolymers. Non-ionic surfactants have been shown to increase membrane fluidity and<br />

contribute to the inhibition of P-gp efflux, resulting in the enhanced cytotoxicity of anthracyclines<br />

in MDR cells. 125,130 However, Pluronic w cannot increase the intracellular accumulation of P-gp<br />

substrate in MRP-expressing cells, indicating that membrane fluidization on its own is insufficient<br />

to inhibit drug efflux in drug-resistant cells. It was believed that the membrane fluidization effect<br />

works synergistically with the ATP-depletion effect induced <strong>by</strong> Pluronic w copolymers to chemosensitize<br />

MDR cells. 127<br />

Recently, Burt’s group reported on the synthesis of a series of low molecular weight PEG-b-<br />

PCL for inhibition of P-gp efflux in human colon adenocarcinoma (Caco-2). Interestingly, at<br />

concentrations above the CMC, the PEG-b-PCL copolymers were found to increase the accumulation<br />

of P-gp substrate, and little activity was observed at concentrations below the CMC. 111<br />

Nevertheless, the inhibition of P-gp efflux was attributed to the unimers, and the inhibition<br />

mechanisms of the PEG-b-PCL copolymers were said to be similar to that of the Pluronic w<br />

copolymers. 107 These low molecular weight PEG-b-PCL copolymers are a class of materials<br />

with potential applications in the delivery of drugs to MDR cancer cells or the formulation of<br />

drugs requiring improved bioavailability.<br />

17.5 IN VIVO BIOLOGICAL PERFORMANCE OF BLOCK COPOLYMER MICELLES<br />

In order to successfully reach and interact with the tumor cells, the micelle-formulated drug must<br />

first circumvent the MPS. As systemic drug delivery vehicles, block copolymer micelles may serve<br />

as either solubilizers and/or true drug carriers depending on their capability for drug retention and<br />

stability in vivo. Several of the micelle formulations that have been developed for hydrophobic<br />

drugs have been shown to provide a significant increase in their water solubility. 36 The use<br />

of biocompatible and biodegradable copolymers to form these micelles may result in formulations<br />

with reduced toxicity in comparison to formulations formed from some conventional surfactants<br />

or cosolvents. As a result, these copolymers may allow for higher doses of the drug to be administered,<br />

and they result in improved efficacy when compared to a conventional formulation of<br />

the drug.<br />

Micelles that function merely as solubilizers increase the aqueous solubility of the drug, but<br />

they do not retain the drug for prolonged periods of time following systemic administration. In this<br />

way, the pharmacokinetics and biodistribution profile of the drug may remain largely unchanged.<br />

By contrast, micelles may function as true carriers that are able to retain the drug until reaching the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 337<br />

Endothelial cells<br />

Red blood cells<br />

Proteins<br />

Macrophages<br />

target or desired site. 16,99,101 Micelles that function as carriers can result in significant improvements<br />

in the toxicity profile and/or therapeutic efficacy of the encapsulated agents.<br />

The ability of a micelle system to function as a solubilizer or a true carrier depends on the<br />

following two factors: the stability of the micelles in vivo and the drug release profile in the<br />

presence of various blood components. The stability of block copolymer micelles in vivo<br />

depends on both thermodynamic and kinetic components as discussed in Section 17.2.<br />

17.5.1 FATE OF POLYMERIC MICELLES IN VIVO<br />

(a)<br />

(b)<br />

(c)<br />

Tumor interstitium<br />

MPS clearance<br />

FIGURE 17.6 In vivo fate of drug-loaded micelles following administration into the bloodstream. (a) Proteins<br />

can be adsorbed onto the micelle surface, leading to eventual clearance <strong>by</strong> the mononuclear phagocytic system<br />

(MPS). (b) The drug-loaded micelles can also circulate in the bloodstream, accumulate in the tumor interstitium<br />

via the enhanced permeability and retention (EPR) effect, and subsequently release the encapsulated<br />

drug to the proximal tumor cells. (c) In the case where the micelle-encapsulated drug has a high affinity for<br />

blood components, the drugs may be released from the micelle core at an accelerated rate to bind to proteins<br />

that are present in the circulation prior to reaching the tumor.<br />

The MPS serves as one of the major mechanisms for the elimination of nanosized drug delivery<br />

systems such as liposomes, nanoparticles, and micelles. 131–133 Clearance <strong>by</strong> the MPS is first<br />

mediated <strong>by</strong> an array of blood components that interact with the delivery vehicle or other<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


338<br />

foreign bodies present within the systemic circulation, a process termed opsonization. 134–136<br />

Following intravenous injection, the micelles quickly distribute into various tissues and organs<br />

<strong>by</strong> way of the blood and lymphatic system. The primary elimination route for micelles is postulated<br />

to include uptake <strong>by</strong> the liver and spleen followed <strong>by</strong> clearance through the kidneys via the<br />

excretion of urine 35,76,137–139 and/or hepatobiliary excretion in the feces. 131,140 In this process,<br />

various plasma proteins play an important role because they adsorb to the surface of a delivery<br />

vehicle within the first few minutes of exposure, especially if the surface is charged or hydrophobic<br />

(Figure 17.6a). 141 The protein adsorption could cause opsonization, leading to rapid clearance of<br />

both carrier and encapsulated drug <strong>by</strong> the MPS. 10,142,143 As a result, a reduction in the extent of<br />

protein adsorption to delivery vehicles such as micelles has been considered to be one of the key<br />

strategies to increase the circulation lifetime of the vehicles in vivo.<br />

Various efforts to prolong the lifetime of micelles in vivo have been explored, including<br />

optimization of micelle size, size distribution, and surface properties. To this point, the micelles<br />

explored for applications in drug delivery have mostly been formed from copolymers that include<br />

PEG as the hydrophilic block. The presence of PEG at the surface of a delivery vehicle is known to<br />

impart stealth properties. 1,2,10 Good correlations have been found between the length of the PEG<br />

block and the extent of stabilization provided. 137 The degree of stabilization is largely determined<br />

<strong>by</strong> the thickness of the PEG layer. The PEG chains are present in a coil-like conformation on the<br />

surface of the micelles. The thickness of the surface layer (dh) may be calculated from the PEG<br />

block length using the following equation: dhZa 0.84 where a is the number of PEG units. 144 In a<br />

study <strong>by</strong> Gref et al. in Balb/C mice, a significant increase in the circulation lifetime of PEG-b-<br />

PLGA nanospheres was observed when the length of the PEG block was increased from 0 to<br />

20,000 g/mol, and the length of the PLGA block was held constant. The increased thickness of<br />

the protective PEG layer likely provides complete coverage of the more hydrophobic PLGA<br />

components. In the same study, the increase in the PEG length also resulted in a significant<br />

reduction in the extent of liver uptake of the particles that provided further evidence that PEG<br />

could effectively decrease the extent of opsonization and, therefore, limit MPS elimination. 66<br />

For materials to be eliminated <strong>by</strong> glomerular filtration through the kidneys, they must have a<br />

molecular weight of less than 50,000 g/mol. 17 The total molecular weight of a block copolymer<br />

micelle is typically on the order O200,000 g/mol, and the molecular weight of the individual<br />

copolymers ranges between 3000 and 25,000 g/mol. 145 Therefore, it may be postulated that only<br />

the copolymer single chains rather than intact micelles are eliminated <strong>by</strong> renal excretion.<br />

17.5.2 IN VIVO DRUG RELEASE PROFILES<br />

Nanotechnology for Cancer Therapy<br />

Micelles introduced into the circulation are exposed to a vast array of proteins, blood cells, and<br />

other blood components. A high affinity between the drug and major blood components commonly<br />

results in an acceleration of drug release from the micelles in vivo that, in turn, leads to a rapid<br />

elimination of the drug from the body (Figure 17.6c). 16 Studies have shown that the in vivo release<br />

profile of drug from micelles is largely determined <strong>by</strong> the compatibility between the drug and the<br />

core-forming block as well as the affinity between the drug and plasma proteins. 36,120,146,147<br />

Ramaswamy et al. investigated the distribution of paclitaxel in vitro in human plasma samples<br />

following incubation of free drug or drug incorporated in PEG-b-PDLLA micelles with the plasma.<br />

In this study, similar plasma distribution profiles were obtained following incubation of free and<br />

micelle-formulated paclitaxel with lipoprotein deficient and lipoprotein rich fractions. 147 In this<br />

way, this study clearly illustrates the extent to which the in vivo retention of drugs in colloidal<br />

carriers may be determined <strong>by</strong> the drugs’ affinity for plasma components. Therefore, the microenvironment<br />

within the delivery system must have the capability to compete with the various<br />

plasma components in order to achieve prolonged drug retention within the micelles in vivo.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 339<br />

17.5.3 TOXICITY, PHARMACOKINETICS, AND ANTI-CANCER EFFICACY OF DRUGS FORMULATED<br />

IN BLOCK COPOLYMER MICELLES<br />

As previously discussed, various micelle-based formulations have been demonstrated, in both<br />

preclinical and clinical studies, to decrease the toxicity and increase the efficacy of the encapsulated<br />

agents. 16,82,90,148 To date, two of the drugs that have been most commonly formulated in block<br />

copolymer micelles are doxorubicin and paclitaxel (Figure 17.3). A general description and overview<br />

of both the doxorubicin and paclitaxel formulations in clinical development has been given in<br />

Section 17.3. At present, there are two micelle formulations of doxorubicin (i.e., SP1049C and<br />

NK911) and two of paclitaxel (Genexol-PM and NK105) that are in clinical trial development.<br />

17.5.3.1 Doxorubicin Formulations (SP1049C and NK911)<br />

SP1049C includes doxorubicin formulated in L61 and F127 Pluronic w copolymers at a ratio of 1:8<br />

(wt/wt). 26 For doxorubicin administered in the SP1049C formulation at a dose of 10 mg/kg, the<br />

AUC in healthy mice was found to be 14.6 mg h/mL while the maximum concentration (Cmax) in<br />

plasma was 4.47 mg/g. For doxorubicin administered as free drug, the AUC is 7.1 mg h/mL and the<br />

Cmax in plasma is 2.77 mg/g. In this way, the SP1049C formulation provides a 2.1-fold increase in<br />

the AUC.<br />

NK911 is another micelle formulation of doxorubicin and includes both chemically and physically<br />

entrapped doxorubicin within PEG-b-PAsp micelles. In its phase I trial, the administration of<br />

the NK911 formulation at a dose of 50 mg/m 2 was shown to provide a significant increase in the<br />

half-life (t1/2) and the AUC as well as a significant decrease in the steady state volume of distribution<br />

(Vss). Specifically, for doxorubicin administered to humans in NK911, the t1/2 values were<br />

reported to be t1/2,aZ7.5 min, t1/2,bZ2.8 h, and t1/2,gZ64.2 h, and the AUC was found to be<br />

3262.7 ng h/mL, and the Vss was 14.9 L/kg. In comparison to free doxorubicin, the t1/2 values<br />

were reported to be t 1/2,aZ2.4 min, t 1/2,bZ0.8 h, and t 1/2,gZ25.8 h, and the AUC was found to<br />

be 1620.3 ng h/mL, and the V ss was 24 L/kg. 82 In this way, the encapsulation of doxorubicin in the<br />

NK911 formulation increases the AUC and plasma Cmax for this drug <strong>by</strong> 36.4 and 28.6-fold,<br />

respectively. The change in the pharmacokinetic profile of the drug translates into changes in the<br />

efficacy and toxicity of the encapsulated agent. Specifically, the anti-tumor activity was evaluated<br />

in four tumor xenograft models in mice, namely, Colon 26, M5076, P388, and Lu-24. For the<br />

groups of animals receiving treatment with the micelle-formulated drug, a significant percentage of<br />

mice bearing colon 26 and MX-1 tumors were cured, whereas no cures were found for mice who<br />

received the same dose of free doxorubicin. Moreover, in all tumor models, treatment with the free<br />

drug resulted in more severe toxicity than that observed in animals receiving even the highest dose<br />

of NK911.<br />

17.5.3.2 Paclitaxel Formulations (Genexol-PM and NK105)<br />

Similar to doxorubicin, there are two promising formulations of paclitaxel that are under clinical<br />

trial evaluation: Genexol-PM and NK105. The main pharmacokinetic parameters of paclitaxel<br />

administered as Genexol-PM at a dose of 135 mg/m 2 are 5473 ng h/mL for the AUC and 12.7 h<br />

for the t 1/2b; whereas for paclitaxel administered as Taxol w , the AUC is 9307.5 ng h/mL, and the<br />

t1/2Z20.1 h. 90,149 In this way, the Genexol-PM formulation does not improve the pharmacokinetic<br />

profile of paclitaxel in vivo; therefore, the micelles in this formulation function primarily as<br />

solubilizers. In fact, it has been found that the drug is released quite rapidly from the PEG-b-<br />

PDLLA micelles following administration. However, as a result of the low toxicity of the micelles<br />

in comparison to Cremophor EL (the primary excipient in Taxol w ), the advantage of this formulation<br />

is the significant improvement in the maximum tolerated dose (i.e., MTDZ390 mg/m 2 )<br />

when compared to the MTD of paclitaxel administered as Taxol w (i.e., 135–200 mg/m 2 ). 90,149<br />

In this case, the increase in the MTD is attributed to the biocompatibility or non-toxicity of the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


340<br />

PEG-b-PDLLA copolymers. Therefore, an improvement in the anti-cancer efficacy may be<br />

expected because of the higher dose of paclitaxel that may be administered when given as the<br />

Genexol-PM formulation.<br />

NK105 is a micelle formulation of paclitaxel wherein the micelles have been shown to perform<br />

as true carriers. The NK105 micelles are formed from copolymers with PEG as the hydrophilic<br />

block and modified PAsp as the core-forming block. The PAsp block is conjugated with 4-phenyl-<br />

1-butanol as a means to increase the compatibility between the micelle core and paclitaxel. Paclitaxel<br />

administered in the NK105 formulation has been shown to have a much longer circulation<br />

lifetime than paclitaxel administered as Taxol w . Specifically, the AUC for paclitaxel was increased<br />

<strong>by</strong> 50–86-fold and the t 1/2 was increased 4–6 times for the NK105 formulation in comparison to<br />

paclitaxel administered as Taxol w . 99 In addition, the therapeutic efficacy of NK105 and Taxol w<br />

formulations were investigated in an HT-29 colon cancer xenograft model in nude mice. The<br />

NK105 formulation was found to exhibit statistically superior anti-tumor activity in comparison<br />

to Taxol w . The anti-tumor activity of NK105 administered at a paclitaxel-equivalent dose of 25 mg/<br />

kg was comparable to that obtained following the administration of Taxol w at a paclitaxel dose of<br />

100 mg/kg. In addition, all tumors disappeared in the NK105 treatment group following one dose of<br />

100 mg/kg, and all mice remained tumor-free for the 33-day duration of the study. This significant<br />

improvement in the anti-cancer activity of paclitaxel following encapsulation in this micelle system<br />

may be attributed to the ability of the NK105 micelles to function as true drug carriers.<br />

17.6 EVOLUTION OF POLYMERIC MICELLES<br />

Recently, much effort has been devoted to increasing drug accumulation at diseased sites. Advances<br />

in synthetic chemistry and a more comprehensive understanding of biological systems have facilitated<br />

the development of smart micelle systems that are capable of targeting drugs to specific tissues<br />

or cellular organelles. Through chemical modification, conventional micelles may be designed to<br />

deliver to and/or release drugs at a specific site of interest <strong>by</strong> active targeting or<br />

external stimulation.<br />

17.6.1 TRIGGERED STIMULATION OF DRUG ACTIVATION OR RELEASE<br />

Following administration of drug-loaded micelles, external stimuli may be locally applied at the<br />

tumor in order to release or activate the drug. This type of drug delivery mechanism may ensure a<br />

more accurate temporal and spatial accumulation of drugs in the tumor as well as reduce toxicity to<br />

other tissues. In this way, localized drug delivery may be achieved in a non-invasive manner using<br />

external stimuli such as ultrasound, heat, and light.<br />

17.6.1.1 Ultrasonic Irradiation<br />

Nanotechnology for Cancer Therapy<br />

The application of ultrasonic irradiation at tumor sites has been found to be effective in enhancing<br />

the accumulation of drugs in tumor cells and has resulted in improved therapeutic efficacy. The use<br />

of ultrasonic irradiation for micelle-based chemotherapy can be optimized <strong>by</strong> varying many factors<br />

including the sonication frequency, the types of waves applied, the time between drug injection and<br />

the application of ultrasound as well as the physico-chemical properties of the micelles. 150,151<br />

Micelle properties such as the hydrophobicity and state of the micelle core were found to be the<br />

most influential in increasing drug uptake in tumor cells upon ultrasonic irradiation. 79,128,152<br />

To date, several mechanisms have been proposed for enhancing tumor cell uptake of micelleformulated<br />

drugs via ultrasonic irradiation. It had been suggested that ultrasound promotes extravasation<br />

of drug-loaded carriers because of increased vasculature permeability. 79 However, a recent<br />

in vivo study revealed that the application of ultrasound does not enhance extravasation of micelles.<br />

Rather, the enhanced drug accumulation in tumor cells requires initial accumulation of drug-loaded<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 341<br />

micelles in the tumor interstitium via the enhanced permeability and retention (EPR) effect. 150 This<br />

indicates that the ultrasonic effect occurs at the cellular level and could potentially play a role in<br />

micelle–cell interactions. In vitro studies have suggested that ultrasound does not increase the<br />

extent of endocytosis of intact micelles. 119 In fact, ultrasound has been shown to trigger drug<br />

release as well as enhance drug uptake <strong>by</strong> increasing cell membrane permeability. 119,151,152<br />

The application of ultrasound perturbs the drug-loaded micelles, shifting the equilibrium from<br />

micelle-encapsulated drug to free drug. In this way, the release of drugs from the micelles is<br />

triggered. 151,152 Furthermore, ultrasonication can cause reversible cell membrane disruption,<br />

creating temporary pores in the membrane, resulting in enhanced drug diffusion into the cells<br />

following drug release. 119 This mechanism is also effective in treating multi-drug resistant<br />

cancer cells because of the enhanced partitioning of the released drug into the intracellular compartment.<br />

79,118 The percentage of drug released after ultrasonication was found to decrease with<br />

increasing hydrophobic interactions between the drug and the micelle core. 151 The use of Pluronic w<br />

copolymers to form micelles that have liquid-like cores facilitates drug release after ultrasonication.<br />

Nonetheless, following injection, these micelles are very unstable in vivo and may result in reduced<br />

drug accumulation at tumor sites. PEG-conjugated distearoylphosphatidylethanolamine (DSPE)<br />

has been added to Pluronic w micelles to improve micelle stability and was proven to be as effective<br />

as PEG-b-PBLA micelles that have solid-like cores in ultrasound-enhanced chemotherapy. 150<br />

Interestingly, after inactivation of ultrasonication, the released drug is re-encapsulated in the<br />

micelles, and the intact drug-loaded micelles are capable of recirculating until they reach the<br />

tumor again. As a result, more than a 3-fold increase in drug uptake in tumor cells is observed<br />

in vivo, and a reduction in the accumulation of the drug in normal tissues is achieved when<br />

compared to non-sonicated tumors. 150<br />

17.6.1.2 Thermosensitive Systems<br />

In addition to ultrasonication, micelles can also be designed to trigger the release of drugs at the<br />

tumors using heat as an external stimulus. Application of thermo-responsive micelles for the<br />

treatment of solid tumors involves the use of a thermosensitive polymer as the micelle-building<br />

block and the local administration of heat at the tumor site. Thermo-responsive copolymers<br />

including poly(N-isopropylacrylamide-b-D,L-lactide) (PIPAAm-b-PDLLA), poly(N-isopropylacrylamide-b-(butyl<br />

methacrylate)) (PIPAAm-b-PBMA), poly(N-isopropylacrylamide-co-N,Ndimethylacrylamide)-b-poly(D,L-lactide-co-glycolide)<br />

(P(IPAAm-co-DMAAm)-b-PLGA) have<br />

been used for the preparation of thermo-responsive micelles. 122,153–155 These micelles undergo<br />

reversible structural changes that facilitate drug release above the lower critical solution temperature<br />

(LCST) of the thermosensitive polymer. In the case of PIPAAm micelles, dehydration of the<br />

outer shells of the micelles also occurs above the LCST, resulting in the enhanced adsorption of<br />

drug-loaded micelles to cells via increased hydrophobic interactions. 155–157<br />

Okano’s group developed a thermo-responsive, doxorubicin-loaded micelle system with an<br />

LCST of 328C. By toggling the temperature below and above the LCST, the release of doxorubicin<br />

may be switched reversibly between an on and off state. At temperatures above the LCST, doxorubicin<br />

release is accelerated, whereas decreasing the temperature to levels below the LCST results<br />

in minimal drug release. 155 This effect is also influenced <strong>by</strong> the state of the micelle core in addition<br />

to the thermosensitivity of the PIPAAm outer shells. For instance, micelles that are formed from<br />

copolymers with liquid-like cores (e.g., TgZ208C for PBMA) are more sensitive to thermal<br />

triggered drug release than micelles with solid-like cores (e.g., TgZ1008C for polystyrene (PS))<br />

at temperatures above the LCST. As a result, the doxorubicin-loaded PPIAAm-b-PBMA micelles<br />

were three times more cytotoxic than the PPIAAm-b-PS micelles above the LCST mainly as a<br />

result of their distinct doxorubicin release profiles. At temperatures below the LCST, both systems<br />

exhibited almost no cytotoxicity. 153<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


342<br />

For in vivo delivery, the LCST of the thermo-responsive micelles should be several degrees<br />

above physiological temperature (i.e., LCSTO378C). 157 The LCST can be manipulated <strong>by</strong> finetuning<br />

the hydrophilicity of the PIPAAm segment. The addition of hydrophilic groups to the<br />

PIPAAm polymer can elevate the LCST of the micelles formed. 122,155–158 For example, Liu et<br />

al. reported that <strong>by</strong> adding hydrophilic DMAAm to the PIPAAm segment, the P(IPAAm-co-<br />

DMAAm)-b-PLGA micelles have an LCST of approximately 398C. As a result, the doxorubicin-loaded<br />

P(IPAAm-co-DMAAm)-b-PLGA micelles have been shown to be more potent in the 4T1<br />

mouse breast cancer cells at 39.58C (IC50Z3.1 mg/L) than at 378C (IC50Z7.9 mg/L). 122 Therefore,<br />

heat-activated drug release in tumors (above 378C) can reduce the possible toxicity associated<br />

with the drug in healthy tissues (at 378C).<br />

17.6.1.3 Photodynamic Therapy<br />

In photodynamic therapy (PDT), light is used to activate the drug (a photosensitizer) at diseased<br />

sites. Initially, PDT was used for the treatment of superficial tumors such as skin cancer. Recent<br />

improvements in the technology have allowed for the treatment of deep tumors such as lung<br />

cancers. 159 The tumor area is exposed to light of a specific wavelength that activates the drug,<br />

leading to the destruction of tumor cells and proximal tumor vasculature via the production of<br />

reactive oxygen species or <strong>by</strong> other mechanisms. 160,161 In order for the PDT agents to be effective<br />

in destroying tumor cells, it is necessary that there is no aggregation of the drug. However, in<br />

aqueous media, the hydrophobic photosensitizers tend to aggregate, resulting in reduced potency.<br />

Pluronic w micelles were reported to be able to solubilize benzoporphyrin while avoiding drug<br />

aggregation in the micelle core, implying that the potency of the photosensitizer was maintained. 162<br />

In this way, polymeric micelles become an attractive alternative for the solubilization and delivery<br />

of photosensitizers.<br />

Kataoka’s group has developed polyion complex (PIC) micelles based on anionic copolymers<br />

(PEG-b-poly(L-lysine) or PEG-b-PAsp) for the encapsulation of a polycationic photosensitizer.<br />

These micelles are stabilized <strong>by</strong> electrostatic interactions, and the micelles are pH-responsive,<br />

allowing intratumoral drug release or intracellular endosomal escape. 160,162,163 The photosensitizer-loaded<br />

PIC micelles showed no skin phototoxicity in rats upon exposure to broadband visible<br />

light, and severe skin phototoxicity was observed with the administration of Photofrin w (a currently<br />

approved PDT drug formulated in 0.9% sodium chloride). 164 The photosensitizer-loaded PIC<br />

micelles were also shown to exhibit enhanced efficacy and reduced toxicity in the Lewis lung<br />

carcinoma (<strong>LLC</strong>) cell line in comparison to free photosensitizer. 160<br />

It has also been reported that the subcellular localization of the photosensitizer influences its<br />

efficacy. Leroux’s group has developed pH-sensitive aluminum chloride pthalocyanine (AICIPc)loaded<br />

N-isopropylacrylamide copolymer micelles for treatment against EMT-6 tumors in vivo.<br />

The tumor accumulation of the micelle-formulated drug is lower than that of AICIPc dissolved in<br />

Cremophor EL. However, the copolymer micelle formulation has been shown to have enhanced<br />

therapeutic efficacy that is most likely attributed to an increase in the cytoplasmic accumulation of<br />

the drug. 165,166 In many cases, the accumulation of photosensitizer at the mitochondria has been<br />

found to be the most efficacious. 161,167 Therefore, the use of pH-responsive polymeric micelles for<br />

PDT could enhance drug accumulation in the mitochondria following endosomal/lysosomal escape<br />

of the micelles. In the future, the design of polymeric micelles for use in PDT will likely be aimed at<br />

improving tumor and subcellular localization <strong>by</strong> fine-tuning the copolymer composition in order to<br />

achieve active targeting.<br />

17.6.2 ACTIVE TARGETING<br />

Nanotechnology for Cancer Therapy<br />

The phenomenon of active targeting of micellar delivery vehicles loaded with drugs implies that the<br />

micelles can accumulate at the targeted cells or tissues without external stimuli via specific sensing<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 343<br />

FIGURE 17.7 (See color insert following page 522.) Confocal fluorescence microscopy image of epidermal<br />

growth factor receptor (EGFR)-overexpressed breast cancer cells (MDA-MB-468) following incubation with<br />

EGF-conjugated poly(ethylene glycol)-block-poly(d-valerolactone) micelles (as shown in red) for 2 h at 378C.<br />

The cell nuclei are stained with Hoechst 33258 (blue). The EGF-conjugated micelles mainly localized at the<br />

perinuclear region as well as within the nucleus as shown <strong>by</strong> the white arrows.<br />

mechanisms that are integrated into the carriers. The carriers can be engineered <strong>by</strong> means of ligand<br />

coupling or addition of pH-sensitive moieties according to the biological characteristics of the<br />

diseased site. In particular, many cancers are associated with over expression of certain cell surface<br />

receptors and acidosis at the tumor sites.<br />

17.6.2.1 Ligand Coupling<br />

It has been demonstrated that ligand-conjugated drug carriers can selectively and efficiently<br />

accumulate in target cells. 168–172 The addition of biomolecules to the surface of micelles enables<br />

the drug-loaded micelles to enter the cells via receptor-mediated endocytosis as shown in<br />

Figure 17.5. Through the highly specific binding affinity between the ligand–receptor complex,<br />

the attachment of ligands to micelles surface allows for targeted drug delivery to specific cells that<br />

express the corresponding cell surface receptors. In particular, many types of cancer cells express a<br />

higher-than-normal level of cell surface receptors that have binding affinities for growth-dependent<br />

molecules as a result of their rapid growth rate. 173 Ligands such as sugar moieties, 174–177 transferrin,<br />

178–180 folate, 78,95,181 epidermal growth factor (EGF), 115,167,178 and antibody fragments 182<br />

have been attached to the surface of copolymer micelles to target different types of cancers.<br />

Kabanov’s group was one of the first to apply the ligand coupling concept to micellar drug<br />

delivery systems. The authors demonstrated that <strong>by</strong> chemically conjugating a2-glycoprotein, a<br />

ligand that targets brain glial cells, to FITC-containing Pluronic w micelles, the accumulation of<br />

FITC in brain tissues was increased and clearance of FITC <strong>by</strong> the lung was decreased in comparison<br />

to delivery using conventional Pluronic w micelles. 126,183 Recently, another ligand, folate, was<br />

conjugated to PEG-b-PLGA micelle system for targeting folate receptor over expressing<br />

cancers. The doxorubicin-loaded folate-PEG-b-PLGA micelles were found to enhance the intracellular<br />

accumulation of doxorubicin and resulted in a lower IC50 in folate receptor overexpressing<br />

KB epidermal carcinoma cells. In vivo studies revealed that treatment with doxorubicin-loaded<br />

folate-PEG-b-PLGA micelles resulted in a high suppression of tumor growth. 78<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


344<br />

In cases where nuclear targeting is desired, ligands such as EGF and the HIV-1 TAT peptide<br />

that have nuclear translocation properties, may be employed. 6,115,184 Zeng et al. reported on an<br />

EGF-conjugated PEG-b-poly(d-valerolactone) (PEG-b-PVL) micelle system that targets EGFRover<br />

expressing breast cancers. The EGF-PEG-b-PVL micelles were shown to mainly localize in<br />

the perinuclear region and in the nucleus of MDA-MB-468 breast cancer cells as shown in<br />

Figure 17.7. 115 Nuclear targeting is critical for the delivery of anti-cancer drugs and oligonucleotides<br />

whose site of action is located in the nucleus. Wagner’s group has developed EGF- and<br />

transferrin-conjugated PEG-b-poly(ethylenimine) (PEG-b-PEI) micelle systems for gene delivery<br />

to tumor tissues. 178,180,185,186 Molecules such as oligonucleotides are more susceptible to<br />

degradation in the endosomal or lysosomal compartments; therefore, pH-responsive micelles<br />

that are capable of destabilizing the endosomal or lysosomal membranes could potentially<br />

improve the therapeutic effect of these molecules.<br />

17.6.2.2 pH-Responsive Systems<br />

Nanotechnology for Cancer Therapy<br />

The pH where drug release is triggered can be adjusted <strong>by</strong> varying the pH-sensitive moiety<br />

incorporated in the copolymer. 158,187 In this way, the pH phase transition of a micelle system<br />

can be designed to suit targets that have different pH values. For instance, micelles that are<br />

stable at pHw2 and are disrupted at pHO7.2 may be suitable for oral delivery of hydrophobic<br />

drugs. As an example, a 10-fold increase in the bioavailability of fenofibrate formulated in pH-responsive<br />

micelles was reported. 188 In the case of tumor targeting, micelles with a phase transition of<br />

approximately pH 7.0 could be beneficial as tumor tissues are known to be slightly more acidic<br />

(mean pH 7.0) than normal tissues (pH 7.4). 187,189,190 If subcellular targeting is desired, cytoplasmic<br />

accumulation of drugs can be enhanced <strong>by</strong> utilizing micelles that are responsive to pH values<br />

ranging from 5.0 to 6.0 (endosomal/lysosomal pH). 187,191,192<br />

pH-responsive micelles have also been developed <strong>by</strong> conjugating drugs to the hydrophobic<br />

blocks of copolymers using pH-sensitive linkages. For instance, doxorubicin can be conjugated to<br />

PEG-b-P(Asp) copolymers via a hydrazone linkage. The micelles formed from this copolymer were<br />

found to be extracellularly stable, and they provided selective release of doxorubicin in the endosomes<br />

where the pH ranges from 5.0 to 6.0. The released drug was found to first localize in the<br />

cytoplasm and eventually in the nucleus. The pH-responsive doxorubicin micelles were also shown<br />

to be capable of infiltrating avascular multicellular tumor spheroids in a uniform manner. Furthermore,<br />

the micelles were effective in suppressing tumor growth in a murine C26 colon tumor model<br />

and increased the MTD <strong>by</strong> more than 2.5 times because of their secure drug encapsulation. 191,192<br />

Another approach to develop pH-responsive micelles is to disrupt micelle formation in the<br />

acidic environment <strong>by</strong> introducing titrable groups into the copolymer. 187 Leroux’s group introduced<br />

a hydrophilic titrable monomer (methacrylic acid or MAA) into the poly(Nisopropylacrylamide)-co-octadecyl<br />

acrylate (poly(NIPA)-co-ODA) copolymer to form pH-responsive<br />

micelles for PDT. 158 These micelles were found to undergo a pH phase transition and structural<br />

changes in the micelle core at a pH of approximately 5.8. The enhanced cytotoxicity of the AICIPcloaded<br />

poly(NIPA)-co-ODA-co-MAA micelles were found to be highly dependent on the pH-gradient<br />

between the cytoplasm and the endosomal/lysosomal compartments. 158<br />

Bae and co-workers have developed several pH-responsive micelle systems <strong>by</strong> using poly<br />

(L-histidine) (polyHis) as the pH-sensitive moiety. The CMC of the PEG-b-polyHis copolymer<br />

was found to increase as the pH of the media was decreased with an abrupt increase in the CMC<br />

occurring at a pH of 7.2. Therefore, with the appropriate alteration in pH, these micelles become<br />

thermodynamically unstable and dissociate, resulting in the release of their contents. 193 Doxorubicin-loaded<br />

PEG-b-polyHis micelles were shown to result in a significant increase in the in vitro<br />

intracellular accumulation of drug, in vivo tumor localization of drug, and tumor growth inhibition<br />

in the A2780 ovarian tumor model via extracellular pH-dependent targeting of the tumor. 194 The<br />

addition of polyHis to PEG-b-PDLLA further reduces the pH phase transition of the micelles to the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 345<br />

range of 6.6–7.2. These micelles were further modified with ligand coupling to achieve a doubletargeting<br />

effect. 181,190 For example, Lee et al. prepared polyHis-b-PEG-b-PDLLA mixed micelles<br />

with biotin as the targeting ligand. Biotin is protected <strong>by</strong> the PEG shell at pHO7.0 and is exposed<br />

on the surface of micelles when the pH is decreased to 6.5–7.0 (extracellular tumor pH). As the pH<br />

is further reduced to less than 6.5 (endosomal pH), the micelles destabilize, triggering drug release<br />

and disrupting endosomal membranes via the proton-sponge effect. 190 In the future, it is anticipated<br />

that combinations of sensing mechanisms will be used to improve the effectiveness of micellar-drug<br />

targeting to the site of interest.<br />

17.7 FUTURE PERSPECTIVES<br />

The field of block copolymer micelles for drug delivery is rapidly expanding and has great potential<br />

in cancer therapeutics. Although there are four promising polymeric micelle formulations currently<br />

in clinical trial development, many preclinical fundamental studies are on-going in order to fully<br />

exploit the use of block copolymer micelles for drug delivery. The advances in synthetic polymer<br />

science continue to provide new biodegradable and biocompatible polymers that can be tailored and<br />

designed for the challenges that anti-cancer drug delivery continues to face. By enhancing the<br />

compatibility between the copolymer and the drug, block copolymer micelles have emerged as true<br />

carriers rather than solubilizers for pharmaceutical agents. The mechanisms of loading and release of<br />

drugs from micelles are being elucidated and will provide a greater understanding of the<br />

relationships between the physico-chemical properties of micelles and their invivo performance.<br />

In the future, multiple targeting mechanisms may be integrated into one micelle system to fully<br />

exploit the use of polymeric micelle as delivery vehicles for anti-cancer drugs.<br />

ACKNOWLEDGMENTS<br />

The authors are grateful to NSERC for funding their research in the area of block copolymer<br />

micelles for applications in drug delivery. In addition, the authors thank Lorne F. Lambier for a<br />

scholarship to H. Lee, CIHR/Rx&D for a career award to C. Allen, a post-doctoral fellowship to P.<br />

Lim Soo, NSERC for a scholarship to J. Liu, and OGS for a scholarship to M. Butler.<br />

REFERENCES<br />

1. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

Theory to practice, Pharmacological Reviews, 53, 283–318, 2001.<br />

2. Torchilin, V. P., Recent advances with liposomes as pharmaceutical carriers, Nature Reviews Drug<br />

Discovery, 4, 145–160, 2005.<br />

3. Allen, C., Maysinger, D., and Eisenberg, A., Nano-engineering block copolymer aggregates for drug<br />

delivery, Colloids and Surfaces B, 16, 3–27, 1999.<br />

4. Kwon, G. S., Diblock copolymer nanoparticles for drug delivery, Critical Reviews in Therapeutic<br />

Drug Carrier Systems, 15, 481–512, 1998.<br />

5. Kataoka, K., Kwon, G. S., Yokoyama, M., Okano, T., and Sakurai, Y., Block-copolymer micelles as<br />

vehicles for drug delivery, Journal of Controlled Release, 24, 119–132, 1993.<br />

6. Gaucher, G., Dufresne, M. H., Sant, V. P., Kang, N., Maysinger, D., and Leroux, J. C., Block<br />

copolymer micelles: Preparation, characterization and application in drug delivery, Journal of<br />

Controlled Release, 109, 169–188, 2005.<br />

7. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., Pluronic (R) block copolymers as novel<br />

polymer therapeutics for drug and gene delivery, Journal of Controlled Release, 82, 189–212, 2002.<br />

8. Elbert, D. L. and Hubbell, J. A., Surface treatments of polymers for biocompatibility, Annual Review<br />

of Material Science, 26, 365–394, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


346<br />

Nanotechnology for Cancer Therapy<br />

9. Harris, J. M., Ed., Poly(ethylene glycol) chemistry: Biotechnical and biomedical applications, In<br />

Topics in Applied Chemistry, Plenum, New York; London, 385, 1992.<br />

10. Allen, C., Dos Santos, N., Gallagher, R., Chiu, G. N. C., Shu, Y., Li, W. M., Johnstone, S. A., Janoff,<br />

A. S., Mayer, L. D., Webb, M. S., and Bally, M. B., Controlling the physical behavior and biological<br />

performance of liposome formulations through use of surface grafted poly(ethylene glycol),<br />

Bioscience Representatives, 22, 225–250, 2002.<br />

11. Gref, R., Domb, A., Quellec, P., Blunk, T., Muller, R. H., Verbavatz, J. M., and Langer, R., The<br />

controlled intravenous delivery of drugs using peg-coated sterically stabilized nanospheres,<br />

Advanced Drug Delivery Reviews, 16, 215–233, 1995.<br />

12. Hunter, R. J. and White, L. R., Foundations of colloid science, Oxford Science Publications, Clarendon<br />

Press, Oxford [Oxfordshire]; Oxford University Press, New York, p. 2 v. (1089), 1987.<br />

13. Müller, R. H., Colloidal Carriers for Controlled Drug Delivery and Targeting: Modification,<br />

Characterization and In Vivo Distribution, Wissenschaftliche Verlagsgesellschaft, Stuttgart; CRC<br />

Press; Sole distributor for North America CRC Press Inc., Boca Raton, FL, p. 379, 1991.<br />

14. Kato, Y., Watanabe, K., Nakakura, M., Hosokawa, T., Hayakawa, E., and Ito, K., Blood clearance<br />

and tissue distribution of various formulations of alpha-tocopherol injection after intravenous administration,<br />

Chemical and Pharmaceutical Bulletin (Tokyo), 41, 599–604, 1993.<br />

15. Oku, N. and Namba, Y., Long-circulating liposomes, Critical Reviews in Therapeutic Drug Carrier<br />

Systems, 11, 231–270, 1994.<br />

16. Nakanishi, T., Fukushima, S., Okamoto, K., Suzuki, M., Matsumura, Y., Yokoyama, M., Okano, T.,<br />

Sakurai, Y., and Kataoka, K., Development of the polymer micelle carrier system for doxorubicin,<br />

Journal of Controlled Release, 74, 295–302, 2001.<br />

17. Seymour, L. W., Duncan, R., Strohalm, J., and Kopecek, J., Effect of molecular weight (Mw) ofN-<br />

(2-hydroxypropyl)methacrylamide copolymers on body distribution and rate of excretion after<br />

subcutaneous, intraperitoneal, and intravenous administration to rats, Journal of Biomedical<br />

Materials Research, 21, 1341–1358, 1987.<br />

18. Allen, C., Yu, Y., Maysinger, D., and Eisenberg, A., Polycaprolactone-b-poly(ethylene oxide) block<br />

copolymer micelles as a novel drug delivery vehicle for neurotrophic agents FK506 and L-685,818,<br />

Bioconjugate Chemistry, 9, 564–572, 1998.<br />

19. Zhang, L. F. and Eisenberg, A., Formation of crew-cut aggregates of various morphologies<br />

from amphiphilic block copolymers in solution, Polymers for Advanced Technologies, 9,<br />

677–699, 1998.<br />

20. Choucair, A. and Eisenberg, A., Control of amphiphilic block copolymer morphologies using<br />

solution conditions, European Physical Journal E, 10, 37–44, 2003.<br />

21. Discher, D. E. and Eisenberg, A., Polymer vesicles, Science, 297, 967–973, 2002.<br />

22. Lim Soo, P. and Eisenberg, A., Preparation of block copolymer vesicles in solution, Journal of<br />

Polymer Science Part B-Polymer Physics, 42, 923–938, 2004.<br />

23. Kabanov, A. V., Nazarova, I. R., Astafieva, I. V., Batrakova, E. V., Alakhov, V. Y., Yaroslavov,<br />

A. A., and Kabanov, V. A., Micelle formation and solubilization of fluorescent-probes in poly(oxyethylene-b-oxypropylene-b-oxyethylene)<br />

solutions, Macromolecules, 28, 2303–2314, 1995.<br />

24. Wilhelm, M., Zhao, C. L., Wang, Y. C., Xu, R. L., Winnik, M. A., Mura, J. L., Riess, G., and<br />

Croucher, M. D., Polymer micelle formation. 3. Poly(styrene-ethylene oxide) block copolymer<br />

micelle formation in water—a fluorescence probe study, Macromolecules, 24, 1033–1040, 1991.<br />

25. Liu, T. B., Kim, K., Hsiao, B. S., and Chu, B., Regular and irregular micelles formed <strong>by</strong> A LEL<br />

triblock copolymer in aqueous solution, Polymer, 45, 7989–7993, 2004.<br />

26. Alakhov, V., Klinski, E., Li, S. M., Pietrzynski, G., Venne, A., Batrakova, E., Bronitch, T., and<br />

Kabanov, A., Block copolymer-based formulation of doxorubicin. From cell screen to clinical trials,<br />

Colloids and Surfaces B: Biointerfaces, 16, 113–134, 1999.<br />

27. Lee, H., Zeng, F. Q., Dunne, M., and Allen, C., Methoxy poly(ethylene glycol)-block-poly(deltavalerolactone)<br />

copolymer micelles for formulation of hydrophobic drugs, Biomacromolecules, 6,<br />

3119–3128, 2005.<br />

28. Kang, N. and Leroux, J. C., Triblock and star-block copolymers of N-(2-hydroxypropyl) methacrylamide<br />

or N-vinyl-2-pyrrolidone and D,L-lactide: Synthesis and self-assembling properties in water,<br />

Polymer, 45, 8967–8980, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 347<br />

29. Alexandridis, P., Holzwarth, J. F., and Hatton, T. A., Micellization of poly(ethylene oxide)–poly(propylene<br />

oxide)–poly(ethylene oxide) triblock copolymers in aqueous-solutions—thermodynamics<br />

of copolymer association, Macromolecules, 27, 2414–2425, 1994.<br />

30. Yamamoto, Y., Yasugi, K., Harada, A., Nagasaki, Y., and Kataoka, K., Temperature-related change<br />

in the properties relevant to drug delivery of poly(ethylene glycol)–poly(D,L-lactide) block copolymer<br />

micelles in aqueous milieu, Journal of Controlled Release, 82, 359–371, 2002.<br />

31. Kang, N., Perron, M. E., Prud’homme, R. E., Zhang, Y. B., Gaucher, G., and Leroux, J. C., Stereocomplex<br />

block copolymer micelles: Core-shell nanostructures with enhanced stability, Nano Letters,<br />

5, 315–319, 2005.<br />

32. Yokoyama, M., Fukushima, S., Uehara, R., Okamoto, K., Kataoka, K., Sakurai, Y., and Okano, T.,<br />

Characterization of physical entrapment and chemical conjugation of adriamycin in polymeric<br />

micelles and their design for in vivo delivery to a solid tumor, Journal of Controlled Release, 50,<br />

79–92, 1998.<br />

33. Croy, S. R. and Kwon, G. S., The effects of pluronic block copolymers on the aggregation state of<br />

nystatin, Journal of Controlled Release, 95, 161–171, 2004.<br />

34. Batrakova, E., Lee, S., Li, S., Venne, A., Alakhov, V., and Kabanov, A., Fundamental relationships<br />

between the composition of pluronic block copolymers and their hypersensitization effect in MDR<br />

cancer cells, Pharmaceutical Research, 16, 1373–1379, 1999.<br />

35. Burt, H. M., Zhang, X. C., Toleikis, P., Embree, L., and Hunter, W. L., Development of copolymers<br />

of poly(D,L-lactide) and methoxypolyethylene glycol as micellar carriers of paclitaxel, Colloids and<br />

Surfaces B: Biointerfaces, 16, 161–171, 1999.<br />

36. Liu, J., Xiao, Y., and Allen, C., Polymer–drug compatibility: A guide to the development of delivery<br />

systems for the anticancer agent, ellipticine, Journal of Pharmaceutical Sciences, 93, 132–143,<br />

2004.<br />

37. Zhang, X. C., Jackson, J. K., and Burt, H. M., Development of amphiphilic diblock copolymers as<br />

micellar carriers of taxol, International Journal of Pharmaceutics, 132, 195–206, 1996.<br />

38. Yokoyama, M., Opanasopit, P., Okano, T., Kawano, K., and Maitani, Y., Polymer design and<br />

incorporation methods for polymeric micelle carrier system containing water-insoluble anticancer<br />

agent camptothecin, Journal of Drug Targeting, 12, 373–384, 2004.<br />

39. Kohori, F., Yokoyama, M., Sakai, K., and Okano, T., Process design for efficient and controlled drug<br />

incorporation into polymeric micelle carrier systems, Journal of Controlled Release, 78, 155–163,<br />

2002.<br />

40. Jette, K. K., Law, D., Schmitt, E. A., and Kwon, G. S., Preparation and drug loading of poly(ethylene<br />

glycol)-block-poly(epsilon-caprolactone) micelles through the evaporation of a cosolvent azeotrope,<br />

Pharmaceutical Research, 21, 1184–1191, 2004.<br />

41. Shuai, X., Ai, H., Nasongkla, N., Kim, S., and Gao, J., Micellar carriers based on block copolymers<br />

of poly(epsilon-caprolactone) and poly(ethylene glycol) for doxorubicin delivery, Journal of<br />

Controlled Release, 98, 415–426, 2004.<br />

42. Zhao, J. X., Allen, C., and Eisenberg, A., Partitioning of pyrene between “crew cut” block copolymer<br />

micelles and H 2O/DMF solvent mixtures, Macromolecules, 30, 7143–7150, 1997.<br />

43. Sharma, P. K. and Bhatia, S. R., Effect of anti-inflammatories on pluronic (R) F127: Micellar<br />

assembly, gelation and partitioning, International Journal of Pharmaceutics, 278, 361–377, 2004.<br />

44. Barreiro-Iglesias, R., Bromberg, L., Temchenko, M., Hatton, T. A., Concheiro, A., and Alvarez-<br />

Lorenzo, C., Solubilization and stabilization of camptothecin in micellar solutions of pluronic-gpoly(acrylic<br />

acid) copolymers, Journal of Controlled Release, 97, 537–549, 2004.<br />

45. Lim Soo, P., Luo, L., Maysinger, D., and Eisenberg, A., Incorporation and release of hydrophobic<br />

probes in biocampatible polycaprolactone-block-poly(ethylene oxide) micelles: Implications for<br />

drug delivery, Langmuir, 18, 9996–10004, 2002.<br />

46. Letchford, K., Zastre, J., Liggins, R., and Burt, H., Synthesis and micellar characterization of short<br />

block length methoxy poly(ethylene glycol)-block-poly(caprolactone) diblock copolymers, Colloids<br />

and Surfaces B: Biointerfaces, 35, 81–91, 2004.<br />

47. Choucair, A. and Eisenberg, A., Interfacial solubilization of model amphiphilic molecules in block<br />

copolymer micelles, Journal of the American Chemical Society, 125, 11993–12000, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


348<br />

Nanotechnology for Cancer Therapy<br />

48. Goldenberg, M. S., Bruno, L. A., and Rennwantz, E. L., Determination of solubilization sites and<br />

efficiency of water-insoluble agents in ethylene oxide-containing nonionic micelles, Journal of<br />

Colloid and Interface Science, 158, 351–363, 1993.<br />

49. Nagarajan, R., Solubilization of “guest” molecules into polymeric aggregates, Polymers for<br />

Advanced Technologies, 12, 23–43, 2001.<br />

50. Nagarajan, R. and Ganesh, K., Comparison of solubilization of hydrocarbons in (PEO–PPO) diblock<br />

versus (PEO–PPO–PEO) triblock copolymer micelles, Journal of Colloid and Interface Science,<br />

184, 489–499, 1996.<br />

51. Kozlov, M. Y., Melik-Nubarov, N. S., Batrakova, E. V., and Kabanov, A. V., Relationship between<br />

pluronic block copolymer structure, critical micellization concentration and partitioning coefficients<br />

of low molecular mass solutes, Macromolecules, 33, 3305–3313, 2000.<br />

52. Xing, L. and Mattice, W. L., Strong solubilization of small molecules <strong>by</strong> triblock-copolymer<br />

micelles in selective solvents, Macromolecules, 30, 1711–1717, 1997.<br />

53. Teng, Y., Morrison, M. E., Munk, P., Webber, S. E., and Prochazka, K., Release kinetics studies of<br />

aromatic molecules into water from block polymer micelles, Macromolecules, 31, 3578–3587, 1998.<br />

54. Lin, W. J., Juang, L. W., and Lin, C. C., Stability and release performance of a series of pegylated<br />

copolymeric micelles, Pharmaceutical Research, 20, 668–673, 2003.<br />

55. Gadelle, F., Koros, W. J., and Schechter, R. S., Solubilization of aromatic solutes in block copolymers,<br />

Macromolecules, 28, 4883–4892, 1995.<br />

56. Lee, J., Cho, E. C., and Cho, K., Incorporation and release behavior of hydrophobic drug in functionalized<br />

poly(D,L-lactide)-block-poly(ethylene oxide) micelles, Journal of Controlled Release, 94,<br />

323–335, 2004.<br />

57. Lavasanifar, A., Samuel, J., and Kwon, G. S., Micelles self-assembled from poly(ethylene oxide)block-poly(N-hexyl<br />

stearate L-aspartamide) <strong>by</strong> a solvent evaporation method: Effect on the solubilization<br />

and haemolytic activity of amphotericin B, Journal of Controlled Release, 77, 155–160,<br />

2001.<br />

58. Adams, M. L. and Kwon, G. S., Relative aggregation state and hemolytic activity of amphotericin B<br />

encapsulated <strong>by</strong> poly(ethylene oxide)-block-poly(N-hexyl-L-aspartamide)-acyl conjugate micelles:<br />

Effects of acyl chain length, Journal of Controlled Release, 87, 23–32, 2003.<br />

59. Oh, K. T., Bronich, T. K., and Kabanov, A. V., Micellar formulations for drug delivery based on<br />

mixtures of hydrophobic and hydrophilic pluronic((R)) block copolymers, Journal of Controlled<br />

Release, 94, 411–422, 2004.<br />

60. Yokoyama, M., Satoh, A., Sakurai, Y., Okano, T., Matsumura, Y., Kakizoe, T., and Kataoka, K.,<br />

Incorporation of water-insoluble anticancer drug into polymeric micelles and control of their particle<br />

size, Journal of Controlled Release, 55, 219–229, 1998.<br />

61. Bromberg, L. and Magner, E., Release of hydrophobic compounds from micellar solutions of<br />

hydrophobically modified polyelectrolytes, Langmuir, 15, 6792–6798, 1999.<br />

62. Kim, S. Y., Ha, J. C., and Lee, Y. M., Poly(ethylene oxide)–poly(propylene oxide)–poly(ethylene<br />

oxide)/poly(epsilon-caprolactone) (PCL) amphiphilic block copolymeric nanospheres. II. Thermoresponsive<br />

drug release behaviors, Journal of Controlled Release, 65, 345–358, 2000.<br />

63. Kim, S. Y., Kim, J. H., Kim, D., An, J. H., Lee, D. S., and Kim, S. C., Drug-releasing kinetics of<br />

MPEG/PLLA block copolymer micelles with different PLLA block lengths, Journal of Applied<br />

Polymer Science, 82, 2599–2605, 2001.<br />

64. Cho, Y. W., Lee, J., Lee, S. C., Huh, K. M., and Park, K., Hydrotropic agents for study of in vitro<br />

paclitaxel release from polymeric micelles, Journal of Controlled Release, 97, 249–257, 2004.<br />

65. Lim Soo, P., Lovric, J., Davidson, P., Maysinger, D., and Eisenberg, A., Polycaprolactone-blockpoly(ethylene<br />

oxide) micelles: A nanodelivery system for 17beta-estradiol, Molecular Pharmacology,<br />

2, 519–527, 2005.<br />

66. Gref, R., Minamitake, Y., Peracchia, M. T., Trubetskoy, V., Torchilin, V., and Langer, R., Biodegradable<br />

long-circulating polymeric nanospheres, Science, 263, 1600–1603, 1994.<br />

67. Nah, J. W., Jeong, Y. I., and Cho, C. S., Norfloxacin release from polymeric micelle of poly(gammabenzyl<br />

L-glutamate) poly(ethylene oxide) poly(gamma-benzylL-glutamate) block copolymer,<br />

Bulletin of the Korean Chemical Society, 19, 962–967, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 349<br />

68. Ryu, J., Jeong, Y. I., Kim, I. S., Lee, J. H., Nah, J. W., and Kim, S. H., Clonazepam release from coreshell<br />

type nanoparticles of poly(epsilon-caprolactone)/poly(ethylene glycol)/poly(epsilon-caprolactone)<br />

triblock copolymers, International Journal of Pharmaceutics, 200, 231–242, 2000.<br />

69. Kim, S. Y., Shin, I. L. G., Lee, Y. M., Cho, C. S., and Sung, Y. K., Methoxy poly(ethylene glycol)<br />

and epsilon-caprolactone amphiphilic block copolymeric micelle containing indomethacin. II.<br />

Micelle formation and drug release behaviours, Journal of Controlled Release, 51, 13–22, 1998.<br />

70. Nah, J. W., Jeong, Y. I., Cho, C. S., and Kim, S. I., Drug-delivery system based on core-shell-type<br />

nanoparticles composed of poly(gamma-benzyl-L-glutamate) and poly(ethylene oxide), Journal of<br />

Applied Polymer Science, 75, 1115–1126, 2000.<br />

71. Leroux, J. C., Allemann, E., DeJaeghere, F., Doelker, E., and Gurny, R., Biodegradable nanoparticles—from<br />

sustained release formulations to improved site specific drug delivery, Journal of<br />

Controlled Release, 39, 339–350, 1996.<br />

72. Kwon, G. S., Naito, M., Kataoka, K., Yokoyama, M., Sakurai, Y., and Okano, T., Block copolymer<br />

micelles as vehicles for hydrophobic drugs, Colloids and Surfaces B: Biointerfaces, 2, 429–434,<br />

1994.<br />

73. Saltzman, W. M., Drug delivery: Engineering principles for drug delivery, Topics in Chemical<br />

Engineering, Oxford University Press, New York, 372, 2001.<br />

74. Doxorubicin, U.S. National Library of Medicine and National Institutes of Health http://www.nlm.<br />

nih.gov/medlineplus/druginfo/medmaster/a682221.html.<br />

75. Lipp, H.-P. and Bokemeyer, C., Anticancer drug toxicity, In Anthracyclines and Other Intercalating<br />

Agents, Lipp, H. P., Ed., Marcel Dekker Inc., New York, pp. 81–113, 1999.<br />

76. Kwon, G. S., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Biodistribution of micelleforming<br />

polymer–drug conjugates, Pharmaceutical Research, 10, 970–974, 1993.<br />

77. Yoo, H. S. and Park, T. G., Biodegradable polymeric micelles composed of doxorubicin conjugated<br />

PLGA–PEG block copolymer, Journal of Controlled Release, 70, 63–70, 2001.<br />

78. Yoo, H. S. and Park, T. G., Folate receptor targeted biodegradable polymeric doxorubicin micelles,<br />

Journal of Controlled Release, 96, 273–283, 2004.<br />

79. Rapoport, N., Pitt, W. G., Sun, H., and Nelson, J. L., Drug delivery in polymeric micelles: From<br />

in vitro to in vivo, Journal of Controlled Release, 91, 85–95, 2003.<br />

80. Pruitt, J. D. and Pitt, W. G., Sequestration and ultrasound-induced release of doxorubicin from<br />

stabilized pluronic P105 micelles, Drug Delivery, 9, 253–258, 2002.<br />

81. Danson, S., Ferry, D., Alakhov, V., Margison, J., Kerr, D., Jowle, D., Brampton, M., Halbert, G., and<br />

Ranson, M., Phase I dose escalation and pharmacokinetic study of pluronic polymer-bound doxorubicin<br />

(SP1049C) in patients with advanced cancer, British Journal of Cancer, 90, 2085–2091, 2004.<br />

82. Matsumura, Y., Hamaguchi, T., Ura, T., Muro, K., Yamada, Y., Shimada, Y., Shirao, K., Okusaka,<br />

T., Ueno, H., Ikeda, M., and Watanabe, N., Phase I clinical trial and pharmacokinetic evaluation of<br />

NK911, a micelle-encapsulated doxorubicin, British Journal of Cancer, 91, 1775–1781, 2004.<br />

83. Kwon, G., Suwa, S., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Enhanced tumor<br />

accumulation and prolonged circulation times of micelle-forming poly(ethylene oxide-aspartate)<br />

block copolymer–adriamycin conjugates, Journal of Controlled Release, 29, 17–23, 1994.<br />

84. Dollery, C. T., Therapeutic Drugs, Churchill Livingstone, Edinburgh, 1999.<br />

85. Yokoyama, M., Miyauchi, M., Yamada, N., Okano, T., Sakurai, Y., Kataoka, K., and Inoue, S.,<br />

Polymer micelles as novel drug carrier—adriamycin-conjugated poly(ethylene glycol) poly(aspartic<br />

acid) block copolymer, Journal of Controlled Release, 11, 269–278, 1990.<br />

86. Fukushima, S., Machida, M., Akutsu, T., Skimizu, K., Tanaka, S., Okamoto, K., and Mashiba, H.,<br />

Roles of adriamycin and adriamycin dimer in antitumor activity of the polymeric micelle carrier<br />

system, Colloids and Surfaces B: Biointerfaces, 16, 227–236, 1999.<br />

87. Paclitaxel, RxList Inc., 2005 http://www.rxlist.com/cgi/generic/paclitaxel.htm.<br />

88. Liggins, R. T., Hunter, W. L., and Burt, H. M., Solid-state characterization of paclitaxel, Journal of<br />

Pharmaceutical Sciences, 86, 1458–1463, 1997.<br />

89. Estimated <strong>by</strong> CAChe 6.1 (computer-aided molecular design modeling software).<br />

90. Kim, T. Y., Kim, D. W., Chung, J. Y., Shin, S. G., Kim, S. C., Heo, D. S., Kim, N. K., and Bang,<br />

Y. J., Phase I and pharmacokinetic study of genexol-PM, a cremophor-free, polymeric micelleformulated<br />

paclitaxel, in patients with advanced malignancies, Clinical Cancer Research, 10,<br />

3708–3716, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


350<br />

Nanotechnology for Cancer Therapy<br />

91. Liggins, R. T. and Burt, H. M., Polyether–polyester diblock copolymers for the preparation of<br />

paclitaxel loaded polymeric micelle formulations, Advanced Drug Delivery Reviews, 54,<br />

191–202, 2002.<br />

92. Le Garrec, D., Gori, S., Luo, L., Lessard, D., Smith, D. C., Yessine, M. A., Ranger, M., and Leroux,<br />

J. C., Poly(N-vinylpyrrolidone)-block-poly(D,L-lactide) as a new polymeric solubilizer for hydrophobic<br />

anticancer drugs: In vitro and in vivo evaluation, Journal of Controlled Release, 99, 83–101,<br />

2004.<br />

93. Shuai, X., Merdan, T., Schaper, A. K., Xi, F., and Kissel, T., Core-cross-linked polymeric micelles as<br />

paclitaxel carriers, Bioconjugate Chemistry, 15, 441–448, 2004.<br />

94. Kim, S. Y. and Lee, Y. M., Taxol-loaded block copolymer nanospheres composed of methoxy<br />

poly(ethylene glycol) and poly(epsilon-caprolactone) as novel anticancer drug carriers, Biomaterials,<br />

22, 1697–1704, 2001.<br />

95. Park, E. K., Lee, S. B., and Lee, Y. M., Preparation and characterization of methoxy poly(ethylene<br />

glycol)/poly(epsilon-caprolactone) amphiphilic block copolymeric nanospheres for tumor-specific<br />

folate-mediated targeting of anticancer drugs, Biomaterials, 26, 1053–1061, 2005.<br />

96. Park, S. R., Oh, D. Y., Kim, D. W., Kim, T. Y., Heo, D. S., Bang, Y. J., Kim, N. K., Kang, W.-K.,<br />

Kim, H.-T., Im, S.-A., Kim, J.-H., and Kim, H.-K., A multi-center, late phase II clinical trial of<br />

genexol (paclitaxel) and cisplatin for patients with advanced gastric cancer, Oncology Reports, 12,<br />

1059–1064, 2004.<br />

97. Kim, S. C., Kim, D. W., Shim, Y. H., Bang, J. S., Oh, H. S., Wan Kim, S., and Seo, M. H., In vivo<br />

evaluation of polymeric micellar paclitaxel formulation: Toxicity and efficacy, Journal of Controlled<br />

Release, 72, 191–202, 2001.<br />

98. Yu, J. J., Jeong, Y. I., Shim, Y. H., and Lim, G. T., Preparation of core-shell type nanoparticles of<br />

diblock copolymers of poly(L-lactide)/poly(ethylene glycol) and their characterization in vitro,<br />

Journal of Applied Polymer Science, 85, 2625–2634, 2002.<br />

99. Hamaguchi, T., Matsumura, Y., Suzuki, M., Shimizu, K., Goda, R., Nakamura, I., Nakatomi, I.,<br />

Yokoyama, M., Kataoka, K., and Kakizoe, T., NK105, a paclitaxel-incorporating micellar nanoparticle<br />

formulation, can extend in vivo antitumour activity and reduce the neurotoxicity of paclitaxel,<br />

British Journal of Cancer, 92, 1240–1246, 2005.<br />

100. Watanabe, M., Kawano, K., Yokoyama, M., Opanasopit, P., Okano, T., and Maitani, Y., Preparation<br />

of camptothecin-loaded polymeric micelles and evaluation of their incorporation and circulation<br />

stability, International Journal of Pharmaceutics, 308, 183–189, <strong>2006</strong>.<br />

101. Nishiyama, N. and Kataoka, K., Preparation and characterization of size-controlled polymeric<br />

micelle containing cis-dichlorodiammineplatinum(II) in the core, Journal of Controlled Release,<br />

74, 83–94, 2001.<br />

102. Uchino, H., Matsumura, Y., Negishi, T., Koizumi, F., Hayashi, T., Honda, T., Nishiyama, N., Kataoka,<br />

K., Naito, S., and Kakizoe, T., Cisplatin-incorporating polymeric micelles (NC-6004) can reduce<br />

nephrotoxicity and neurotoxicity of cisplatin in rats, British Journal of Cancer, 93, 678–687, 2005.<br />

103. Maysinger, D., Berezovska, O., Savic, R., Lim Soo, P., and Eisenberg, A., Block copolymers modify<br />

the internalization of micelle-incorporated probes into neural cells, Biochimica et Biophysica Acta,<br />

1539, 205–217, 2001.<br />

104. Allen, C., Yu, Y., Eisenberg, A., and Maysinger, D., Cellular internalization of PCL(20)-b-PEO(44)<br />

block copolymer micelles, Biochimica et Biophysica Acta, 1421, 32–38, 1999.<br />

105. Savic, R., Luo, L., Eisenberg, A., and Maysinger, D., Micellar nanocontainers distribute to defined<br />

cytoplasmic organelles, Science, 300, 615–618, 2003.<br />

106. Luo, L., Tam, J., Maysinger, D., and Eisenberg, A., Cellular internalization of poly(ethylene oxide)b-poly(epsilon-caprolactone)<br />

diblock copolymer micelles, Bioconjugate Chemistry, 13, 1259–1265,<br />

2002.<br />

107. Zastre, J., Jackson, J., and Burt, H., Evidence for modulation of P-glycoprotein-mediated efflux <strong>by</strong><br />

methoxypolyethylene glycol-block-polycaprolactone amphiphilic diblock copolymers, Pharmaceutical<br />

Research, 21, 1489–1497, 2004.<br />

108. Alakhov, V., Moskaleva, E., Batrakova, E. V., and Kabanov, A. V., Hypersensitization of multidrug<br />

resistant human ovarian carcinoma cells <strong>by</strong> pluronic P85 block copolymer, Bioconjugate Chemistry,<br />

7, 209–216, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 351<br />

109. Batrakova, E. V., Li, S., Elmquist, W. F., Miller, D. W., Alakhov, V. Y., and Kabanov, A. V.,<br />

Mechanism of sensitization of MDR cancer cells <strong>by</strong> pluronic block copolymers: Selective energy<br />

depletion, British Journal of Cancer, 85, 1987–1997, 2001.<br />

110. Batrakova, E. V., Han, H. Y., Alakhov, V., Miller, D. W., and Kabanov, A. V., Effects of pluronic<br />

block copolymers on drug absorption in Caco-2 cell monolayers, Pharmaceutical Research, 15,<br />

850–855, 1998.<br />

111. Zastre, J., Jackson, J., Bajwa, M., Liggins, R., Iqbal, F., and Burt, H., Enhanced cellular accumulation<br />

of a P-glycoprotein substrate, rhodamine-123, <strong>by</strong> Caco-2 cells using low molecular weight<br />

methoxypolyethylene glycol-block-polycaprolactone diblock copolymers, European Journal of<br />

Pharmaceutics and Biopharmaceutics, 54, 299–309, 2002.<br />

112. Robinson, J. R. and Lee, V. H. L., Eds., Controlled drug delivery: Fundamentals and applications,<br />

Drugs and the Pharmaceutical Sciences, Vol. 29, Marcel Dekker, New York, pp. xix, 716, 1987.<br />

113. Pastan, I. H. and Willingham, M. C., Eds., Endocytosis, Plenum Press, New York, pp. xviii, 326,<br />

1985.<br />

114. Torchilin, V. P., Targeted polymeric micelles for delivery of poorly soluble drugs, Cellular and<br />

Molecular Life Sciences, 61, 2549–2559, 2004.<br />

115. Zeng, F., Lee, H., and Allen, C., Epidermal growth factor-conjugated poly(ethylene glycol)-blockpoly(d-valerolactone)<br />

copolymer micelles for targeted delivery of chemotherapeutics, Bioconjugate<br />

Chemistry, <strong>2006</strong>, in Press.<br />

116. Kakizawa, Y. and Kataoka, K., Block copolymer micelles for delivery of gene and related<br />

compounds, Advanced Drug Delivery Reviews, 54, 203–222, 2002.<br />

117. Muniruzzaman, M., Marin, A., Luo, Y., Prestwich, G. D., Pitt, W. G., Husseini, G., and Rapoport,<br />

N. Y., Intracellular uptake of pluronic copolymer: Effects of the aggregation state, Colloids and<br />

Surfaces B: Biointerfaces, 25, 233–241, 2002.<br />

118. Rapoport, N., Marin, A., Luo, Y., Prestwich, G. D., and Muniruzzaman, M. D., Intracellular uptake<br />

and trafficking of pluronic micelles in drug-sensitive and MDR cells: Effect on the intracellular drug<br />

localization, Journal of Pharmaceutical Science, 91, 157–170, 2002.<br />

119. Husseini, G. A., Runyan, C. M., and Pitt, W. G., Investigating the mechanism of acoustically<br />

activated uptake of drugs from pluronic micelles, BMC Cancer, 2, 20, 2002.<br />

120. Liu, J., Zeng, F., and Allen, C., Influence of serum protein on polycarbonate-based copolymer<br />

micelles as a delivery system for a hydrophobic anti-cancer agent, Journal of Controlled Release,<br />

103, 481–497, 2005.<br />

121. Pante, N. and Kann, M., Nuclear pore complex is able to transport macromolecules with diameters of<br />

about 39 nm, Molecular Biology of the Cell, 13, 425–434, 2002.<br />

122. Liu, S. Q., Tong, Y. W., and Yang, Y. Y., Incorporation and in vitro release of doxorubicin in<br />

thermally sensitive micelles made from poly(N-isopropylacrylamide-co-N,N-dimethylacrylamide)b-poly(D,L-lactide-co-glycolide)<br />

with varying compositions, Biomaterials, 26, 5064–5074, 2005.<br />

123. Kabanov, A. V., Batrakova, E. V., and Miller, D. W., Pluronic block copolymers as modulators of<br />

drug efflux transporter activity in the blood–brain barrier, Advanced Drug Delivery Reviews, 55,<br />

151–164, 2003.<br />

124. Batrakova, E. V., Li, S., Miller, D. W., and Kabanov, A. V., Pluronic P85 increases permeability of a<br />

broad spectrum of drugs in polarized BBMEC and Caco-2 cell monolayers, Pharmaceutical<br />

Research, 16, 1366–1372, 1999.<br />

125. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., Pluronic block copolymers for overcoming<br />

drug resistance in cancer, Advanced Drug Delivery Reviews, 54, 759–779, 2002.<br />

126. Kabanov, A. V., Batrakova, E. V., Meliknubarov, N. S., Fedoseev, N. A., Dorodnich, T. Y., Alakhov,<br />

V. Y., Chekhonin, V. P., Nazarova, I. R., and Kabanov, V. A., A new class of drug carriers—micelles<br />

of poly(oxyethylene)–poly(oxypropylene) block copolymers as microcontainers for drug targeting<br />

from blood in brain, Journal of Controlled Release, 22, 141–157, 1992.<br />

127. Miller, D. W., Batrakova, E. V., and Kabanov, A. V., Inhibition of multidrug resistance-associated<br />

protein (MRP) functional activity with pluronic block copolymers, Pharmaceutical Research, 16,<br />

396–401, 1999.<br />

128. Rapoport, N., Marin, A. P., and Timoshin, A. A., Effect of a polymeric surfactant on electron<br />

transport in HL-60 cells, Archives of Biochemistry and Biophysics, 384, 100–108, 2000.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


352<br />

Nanotechnology for Cancer Therapy<br />

129. Slepnev, V. I., Kuznetsova, L. E., Gubin, A. N., Batrakova, E. V., Alakhov, V., and Kabanov, A. V.,<br />

Micelles of poly(oxyethylene)–poly(oxypropylene) block copolymer (pluronic) as a tool for lowmolecular<br />

compound delivery into a cell: Phosphorylation of intracellular proteins with micelle<br />

incorporated [gamma-32P]ATP, Biochemistry International, 26, 587–595, 1992.<br />

130. Regev, R., Assaraf, Y. G., and Eytan, G. D., Membrane fluidization <strong>by</strong> ether, other anesthetics, and<br />

certain agents abolishes P-glycoprotein ATPase activity and modulates efflux from multidrug-resistant<br />

cells, European Journal of Biochemistry, 259, 18–24, 1999.<br />

131. Yamamoto, Y., Nagasaki, Y., Kato, Y., Sugiyama, Y., and Kataoka, K., Long-circulating poly(ethylene<br />

glycol)–poly(D,L-lactide) block copolymer micelles with modulated surface charge, Journal<br />

of Controlled Release, 77, 27–38, 2001.<br />

132. Poznansky, M. J. and Juliano, R. L., Biological approaches to the controlled delivery of drugs: A<br />

critical review, Pharmacological Reviews, 36, 277–336, 1984.<br />

133. Drummond, D. C., Meyer, O., Hong, K., Kirpotin, D. B., and Papahadjopoulos, D., Optimizing<br />

liposomes for delivery of chemotherapeutic agents to solid tumors, Pharmacological Reviews, 51,<br />

691–743, 1999.<br />

134. Patel, H. M., Serum opsonins and liposomes: Their interaction and opsonophagocytosis, Critical<br />

Reviews in Therapeutic Drug Carrier Systems, 9, 39–90, 1992.<br />

135. Moghimi, S. M. and Patel, H. M., Tissue specific opsonins for phagocytic cells and their different<br />

affinity for cholesterol-rich liposomes, FEBS Letters, 233, 143–147, 1988.<br />

136. Ulrich, F. and Zilversmit, D. B., Release from alveolar macrophages of an inhibitor of phagocytosis,<br />

American Journal of Physiology, 218, 1118–1127, 1970.<br />

137. Dunn, S. E., Brindley, A., Davis, S. S., Davies, M. C., and Illum, L., Polystyrene–poly(ethylene<br />

glycol) (PS–PEG2000) particles as model systems for site specific drug delivery. 2. The effect of<br />

PEG surface density on the in vitro cell interaction and in vivo biodistribution, Pharmaceutical<br />

Research, 11, 1016–1022, 1994.<br />

138. Batrakova, E. V., Li, S., Li, Y., Alakhov, V. Y., Elmquist, W. F., and Kabanov, A. V., Distribution<br />

kinetics of a micelle-forming block copolymer pluronic P85, Journal of Controlled Release, 100,<br />

389–397, 2004.<br />

139. Grindel, J. M., Jaworski, T., Piraner, O., Emanuele, R. M., and Balasubramanian, M., Distribution,<br />

metabolism, and excretion of a novel surface-active agent, purified poloxamer 188, in rats, dogs, and<br />

humans, Journal of Pharmaceutical Science, 91, 1936–1947, 2002.<br />

140. Novakova, K., Laznicek, M., Rypacek, F., and Machova, L., I-125-labeled PLA/PEO block copolymer:<br />

Biodistribution studies in rats, Journal of Bioactive and Compatible Polymers, 17, 285–296,<br />

2002.<br />

141. Ishihara, K., Nomura, H., Mihara, T., Kurita, K., Iwasaki, Y., and Nakabayashi, N., Why do phospholipid<br />

polymers reduce protein adsorption? Journal of Biomedical Materials Research, 39,<br />

323–330, 1998.<br />

142. Devine, D. V. and Marjan, J. M., The role of immunoproteins in the survival of liposomes in the<br />

circulation, Critical Reviews in Therapeutic Drug Carrier Systems, 14, 105–131, 1997.<br />

143. Patel, H. M. and Moghimi, S. M., Serum-mediated recognition of liposomes <strong>by</strong> phagocytic cells of<br />

the reticuloendothelial system—the concept of tissue specificity, Advanced Drug Delivery Reviews,<br />

32, 45–60, 1998.<br />

144. Tan, J. S., Butterfield, D. E., Voycheck, C. L., Caldwell, K. D., and Li, J. T., Surface modification of<br />

nanoparticles <strong>by</strong> PEO/PPO block copolymers to minimize interactions with blood components and<br />

prolong blood circulation in rats, Biomaterials, 14, 823–833, 1993.<br />

145. Allen, C., Maysinger, D., and Eisenberg, A., Nano-engineering block copolymer aggregates for drug<br />

delivery, Colloids and Surfaces B: Biointerfaces, 16, 3–27, 1999.<br />

146. Seki, J., Sonoke, S., Saheki, A., Koike, T., Fukui, H., Doi, M., and Mayumi, T., Lipid transfer protein<br />

transports compounds from lipid nanoparticles to plasma lipoproteins, International Journal of<br />

Pharmaceutics, 275, 239–248, 2004.<br />

147. Ramaswamy, M., Zhang, X., Burt, H. M., and Wasan, K. M., Human plasma distribution of free<br />

paclitaxel and paclitaxel associated with diblock copolymers, Journal of Pharmaceutical Science,<br />

86, 460–464, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 353<br />

148. Mizumura, Y., Matsumura, Y., Hamaguchi, T., Nishiyama, N., Kataoka, K., Kawaguchi, T.,<br />

Hrushesky, W. J., Moriyasu, F., and Kakizoe, T., Cisplatin-incorporated polymeric micelles eliminate<br />

nephrotoxicity, while maintaining antitumor activity, Japan Journal of Cancer Research, 92,<br />

328–336, 2001.<br />

149. Gianni, L., Kearns, C. M., Giani, A., Capri, G., Vigano, L., Lacatelli, A., Bonadonna, G., and Egorin,<br />

M. J., Nonlinear pharmacokinetics and metabolism of paclitaxel and its pharmacokinetic/pharmacodynamic<br />

relationships in humans, Journal of Clinical Oncology, 13, 180–190, 1995.<br />

150. Gao, Z. G., Fain, H. D., and Rapoport, N., Controlled and targeted tumor chemotherapy <strong>by</strong> micellarencapsulated<br />

drug and ultrasound, Journal of Controlled Release, 102, 203–222, 2005.<br />

151. Husseini, G. A., Myrup, G. D., Pitt, W. G., Christensen, D. A., and Rapoport, N. Y., Factors affecting<br />

acoustically triggered release of drugs from polymeric micelles, Journal of Controlled Release, 69,<br />

43–52, 2000.<br />

152. Marin, A., Muniruzzaman, M., and Rapoport, N., Mechanism of the ultrasonic activation of micellar<br />

drug delivery, Journal of Controlled Release, 75, 69–81, 2001.<br />

153. Chung, J. E., Yokoyama, M., and Okano, T., Inner core segment design for drug delivery control of<br />

thermo-responsive polymeric micelles, Journal of Controlled Release, 65, 93–103, 2000.<br />

154. Kohori, F., Sakai, K., Aoyagi, T., Yokoyama, M., Sakurai, Y., and Okano, T., Preparation and<br />

characterization of thermally responsive block copolymer micelles comprising poly(N-isopropylacrylamide-b-DL-lactide),<br />

Journal of Controlled Release, 55, 87–98, 1998.<br />

155. Chung, J. E., Yokoyama, M., Yamato, M., Aoyagi, T., Sakurai, Y., and Okano, T., Thermo-responsive<br />

drug delivery from polymeric micelles constructed using block copolymers of poly(Nisopropylacrylamide)<br />

and poly(butylmethacrylate), Journal of Controlled Release, 62, 115–127,<br />

1999.<br />

156. Chung, J. E., Yokoyama, M., Aoyagi, T., Sakurai, Y., and Okano, T., Effect of molecular architecture<br />

of hydrophobically modified poly(N-isopropylacrylamide) on the formation of thermoresponsive<br />

core-shell micellar drug carriers, Journal of Controlled Release, 53, 119–130, 1998.<br />

157. Cammas, S., Suzuki, K., Sone, C., Sakurai, Y., Kataoka, K., and Okano, T., Thermo-responsive<br />

polymer nanoparticles with a core-shell micelle structure as site-specific drug carriers, Journal of<br />

Controlled Release, 48, 157–164, 1997.<br />

158. Taillefer, J., Jones, M. C., Brasseur, N., van Lier, J. E., and Leroux, J. C., Preparation and characterization<br />

of pH-responsive polymeric micelles for the delivery of photosensitizing anticancer drugs,<br />

Journal of Pharmaceutical Science, 89, 52–62, 2000.<br />

159. Houlton, S., Blocking the way forward, 2003 http://www.users.globalnet.uk/~sarahx/articles/<br />

mcblocks.htm.<br />

160. Zhang, G. D., Harada, A., Nishiyama, N., Jiang, D. L., Koyama, H., Aida, T., and Kataoka, K.,<br />

Polyion complex micelles entrapping cationic dendrimer porphyrin: Effective photosensitizer for<br />

photodynamic therapy of cancer, Journal of Controlled Release, 93, 141–150, 2003.<br />

161. Dolmans, D. E., Fukumura, D., and Jain, R. K., Photodynamic therapy for cancer, Nature Reviews<br />

Cancer, 3, 380–387, 2003.<br />

162. van Nostrum, C. F., Polymeric micelles to deliver photosensitizers for photodynamic therapy,<br />

Advanced Drug Delivery Reviews, 56, 9–16, 2004.<br />

163. Jang, W. D., Nishiyama, N., Zhang, G. D., Harada, A., Jiang, D. L., Kawauchi, S., Morimoto, Y.,<br />

Kikuchi, M., Koyama, H., Aida, T., and Kataoka, K., Supramolecular nanocarrier of anionic<br />

dendrimer porphyrins with cationic block copolymers modified with polyethylene glycol to<br />

enhance intracellular photodynamic efficacy, Angewandte Chemie International Edition in<br />

English, 44, 419–423, 2005.<br />

164. Ideta, R., Tasaka, F., Jang, W. D., Nishiyama, N., Zhang, G. D., Harada, A., Yanagi, Y., Tamaki, Y.,<br />

Aida, T., and Kataoka, K., Nanotechnology-based photodynamic therapy for neovascular disease<br />

using a supramolecular nanocarrier loaded with a dendritic photosensitizer, Nano Letters, 5,<br />

2426–2431, 2005.<br />

165. Taillefer, J., Brasseur, N., van Lier, J. E., Lenaerts, V., Le Garrec, D., and Leroux, J. C., In-vitro and<br />

in-vivo evaluation of pH-responsive polymeric micelles in a photodynamic cancer therapy model,<br />

Journal of Pharmacy and Pharmacology, 53, 155–166, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


354<br />

Nanotechnology for Cancer Therapy<br />

166. Le Garrec, D., Taillefer, J., Van Lier, J. E., Lenaerts, V., and Leroux, J. C., Optimizing pH-responsive<br />

polymeric micelles for drug delivery in a cancer photodynamic therapy model, Journal of Drug<br />

Targeting, 10, 429–437, 2002.<br />

167. Woodburn, K. W., Vardaxis, N. J., Hill, J. S., Kaye, A. H., Reiss, J. A., and Phillips, D. R., Evaluation<br />

of porphyrin characteristics required for photodynamic therapy, Photochemistry and Photobiology,<br />

55, 697–704, 1992.<br />

168. Mamot, C., Drummond, D. C., Greiser, U., Hong, K., Kirpotin, D. B., Marks, J. D., and Park, J. W.,<br />

Epidermal growth factor receptor (EGFR)-targeted immunoliposomes mediate specific and efficient<br />

drug delivery to EGFR- and EGFRvIII-overexpressing tumor cells, Cancer Research, 63,<br />

3154–3161, 2003.<br />

169. Torchilin, V. P., Lukyanov, A. N., Gao, Z. G., and Papahadjopoulos-Sternberg, B., Immunomicelles:<br />

Targeted pharmaceutical carriers for poorly soluble drugs, Proceedings of the National Academy of<br />

Sciences of the United States of America, 100, 6039–6044, 2003.<br />

170. Torchilin, V. P., Fluorescence microscopy to follow the targeting of liposomes and micelles to cells<br />

and their intracellular fate, Advanced Drug Delivery Reviews, 57, 95–109, 2005.<br />

171. Farokhzad, O. C., Jon, S. Y., Khadelmhosseini, A., Tran, T. N. T., LaVan, D. A., and Langer, R.,<br />

Nanopartide-aptamer bioconjugates: A new approach for targeting prostate cancer cells, Cancer<br />

Research, 64, 7668–7672, 2004.<br />

172. Lee, H., Jang, I. H., Ryu, S. H., and Park, T. G., N-terminal site-specific mono-PEGylation of<br />

epidermal growth factor, Pharmaceutical Research, 20, 818–825, 2003.<br />

173. Mendelsohn, J. and Baselga, J., The EGF receptor family as targets for cancer therapy, Oncogene,<br />

19, 6550–6565, 2000.<br />

174. Wakebayashi, D., Nishiyama, N., Yamasaki, Y., Itaka, K., Kanayama, N., Harada, A., Nagasaki, Y.,<br />

and Kataoka, K., Lactose-conjugated polyion complex micelles incorporating plasmid DNA as a<br />

targetable gene vector system: Their preparation and gene transfecting efficiency against cultured<br />

HepG2 cells, Journal of Controlled Release, 95, 653–664, 2004.<br />

175. Yasugi, K., Nakamura, T., Nagasaki, Y., Kato, M., and Kataoka, K., Sugar-installed polymer<br />

micelles: Synthesis and micellization of poly(ethylene glycol)–poly(D,L-lactide) block copolymers<br />

having sugar groups at the PEG chain end, Macromolecules, 32, 8024–8032, 1999.<br />

176. Jule, E., Nagasaki, Y., and Kataoka, K., Lactose-installed poly(ethylene glycol)–poly(D,L-lactide)<br />

block copolymer micelles exhibit fast-rate binding and high affinity toward a protein bed simulating<br />

a cell surface. A surface plasmon resonance study, Bioconjugate Chemistry, 14, 177–186, 2003.<br />

177. Lim, D. W., Yeom, Y. I., and Park, T. G., Poly(DMAEMA-NVP)-b-PEG-galactose as gene delivery<br />

vector for hepatocytes, Bioconjugate Chemistry, 11, 688–695, 2000.<br />

178. Kursa, M., Walker, G. F., Roessler, V., Ogris, M., Roedl, W., Kircheis, R., and Wagner, E., Novel<br />

shielded transferrin–polyethylene glycol–polyethylenimine/DNA complexes for systemic tumortargeted<br />

gene transfer, Bioconjugate Chemistry, 14, 222–231, 2003.<br />

179. Vinogradov, S., Batrakova, E., Li, S., and Kabanov, A., Polyion complex micelles with proteinmodified<br />

corona for receptor-mediated delivery of oligonucleotides into cells, Bioconjugate Chemistry,<br />

10, 851–860, 1999.<br />

180. Ogris, M., Walker, G., Blessing, T., Kircheis, R., Wolschek, M., and Wagner, E., Tumor-targeted<br />

gene therapy: Strategies for the preparation of ligand-polyethylene glycol-polyethylenimine/DNA<br />

complexes, Journal of Controlled Release, 91, 173–181, 2003.<br />

181. Lee, E. S., Na, K., and Bae, Y. H., Polymeric micelle for tumor pH and folate-mediated targeting,<br />

Journal of Controlled Release, 91, 103–113, 2003.<br />

182. Merdan, T., Callahan, J., Peterson, H., Bakowsky, U., Kopeckova, P., Kissel, T., and Kopecek, J.,<br />

Pegylated polyethylenimine-Fab 0 antibody fragment conjugates for targeted gene delivery to human<br />

ovarian carcinoma cells, Bioconjugate Chemistry, 14, 989–996, 2003.<br />

183. Kabanov, A. V. and Alakhov, V. Y., Amphiphilic Block Copolymers: Self-Assemble and Applications<br />

In Micelles of Amphiphilic Block Copolymers as Vehicles for Drug Delivery, Alexandris, P.<br />

and Lindman, B., Eds., Elsevier, The Netherlands, 1997.<br />

184. Lin, S. Y., Makino, K., Xia, W. Y., Matin, A., Wen, Y., Kwong, K. Y., Bourguignon, L., and Hung,<br />

M. C., Nuclear localization of EGF receptor and its potential new role as a transcription factor,<br />

Nature Cell Biology, 3, 802–808, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles for Formulation of Anti-Cancer Drugs 355<br />

185. Kircheis, R., Kichler, A., Wallner, G., Kursa, M., Ogris, M., Felzmann, T., Buchberger, M., and<br />

Wagner, E., Coupling of cell-binding ligands to polyethylenimine for targeted gene delivery, Gene<br />

Therapy, 4, 409–418, 1997.<br />

186. Blessing, T., Kursa, M., Holzhauser, R., Kircheis, R., and Wagner, E., Different strategies for<br />

formation of PEGylated EGF-conjugated PEI/DNA complexes for targeted gene delivery, Bioconjugate<br />

Chemistry, 12, 529–537, 2001.<br />

187. Gillies, E. R. and Frechet, J. M. J., Development of acid-sensitive copolymer micelles for drug<br />

delivery, Pure and Applied Chemistry, 76, 1295–1307, 2004.<br />

188. Sant, V. P., Smith, D., and Leroux, J. C., Enhancement of oral bioavailability of poorly water-soluble<br />

drugs <strong>by</strong> poly(ethylene glycol)-block-poly(alkyl acrylate-co-methacrylic acid) self-assemblies,<br />

Journal of Controlled Release, 104, 289–300, 2005.<br />

189. Tannock, I. F. and Rotin, D., Acid pH in tumors and its potential for therapeutic exploitation, Cancer<br />

Research, 49, 4373–4384, 1989.<br />

190. Lee, E. S., Na, K., and Bae, Y. H., Super pH-sensitive multifunctional polymeric micelle, Nano<br />

Letters, 5, 325–329, 2005.<br />

191. Bae, Y., Fukushima, S., Harada, A., and Kataoka, K., Design of environment-sensitive supramolecular<br />

assemblies for intracellular drug delivery: Polymeric micelles that are responsive to<br />

intracellular pH change, Angewandte Chemie International Edition in English, 42, 4640–4643, 2003.<br />

192. Bae, Y., Nishiyama, N., Fukushima, S., Koyama, H., Yasuhiro, M., and Kataoka, K., Preparation and<br />

biological characterization of polymeric micelle drug carriers with intracellular pH-triggered drug<br />

release property: Tumor permeability, controlled subcellular drug distribution, and enhanced in vivo<br />

antitumor efficacy, Bioconjugate Chemistry, 16, 122–130, 2005.<br />

193. Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., Poly(L-histidine)–PEG block copolymer micelles and<br />

pH-induced destabilization, Journal of Controlled Release, 90, 363–374, 2003.<br />

194. Gao, Z. G., Lee, D. H., Kim, D. I., and Bae, Y. H., Doxorubicin loaded pH-sensitive micelle targeting<br />

acidic extracellular pH of human ovarian A2780 tumor in mice, Journal of Drug Targeting, 13,<br />

391–397, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


18 PEO-Modified<br />

Poly(L-Amino Acid) Micelles<br />

for Drug Delivery<br />

CONTENTS<br />

Xiao-Bing Xiong, Hamidreza Montazeri Aliabadi, and<br />

Afsaneh Lavasanifar<br />

18.1 Introduction ....................................................................................................................... 358<br />

18.2 Synthesis of PEO-b-PLAA Block Copolymers ................................................................ 358<br />

18.3 Micellization of PEO-b-PLAA Block Copolymers .......................................................... 361<br />

18.3.1 Dialysis Method .................................................................................................. 362<br />

18.3.2 Solvent Evaporation Method .............................................................................. 362<br />

18.3.3 Co-Solvent Evaporation Method ........................................................................ 362<br />

18.4 Rational Design and Functional Properties of PEO-b-PLAA Micelles in<br />

Drug Delivery .................................................................................................................... 363<br />

18.4.1 The PEO Shell..................................................................................................... 363<br />

18.4.2 The PLAA Core .................................................................................................. 363<br />

18.4.3 Micellar Dimensions ........................................................................................... 363<br />

18.4.4 Micellar Stability................................................................................................. 367<br />

18.4.5 Drug Incorporation and Release Properties........................................................ 368<br />

18.5 PEO-b-PLAA Micelles for the Delivery of Anti-Cancer Drugs ...................................... 369<br />

18.5.1 Cyclophosphamide .............................................................................................. 369<br />

18.5.2 Doxorubicin ......................................................................................................... 369<br />

18.5.3 Paclitaxel ............................................................................................................. 370<br />

18.5.4 Cisplatin and Its Derivative ................................................................................ 370<br />

18.5.5 Methotrexate........................................................................................................ 371<br />

18.5.6 KRN5500............................................................................................................. 371<br />

18.5.7 Camptothecin....................................................................................................... 372<br />

18.6 PEO-b-PLAA Micelles for the Delivery of Other Therapeutic Agents........................... 372<br />

18.6.1 Amphotericin B ................................................................................................... 372<br />

18.6.2 Indomethacin ....................................................................................................... 373<br />

18.7 PEO-b-PLAA Micelles for Active Drug Targeting.......................................................... 373<br />

18.8 PEO-b-PLAA Micelles for Gene Delivery ....................................................................... 374<br />

References .................................................................................................................................... 376<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

357


358<br />

18.1 INTRODUCTION<br />

Polymeric micelles are core/shell structures formed through the self-assembly of amphiphilic block<br />

copolymers. The nanoscopic dimension as well as unique properties offered <strong>by</strong> separated core and<br />

shell domains in the structure of polymeric micelles has made them one of the most promising<br />

carriers for passive or active drug targeting in cancer. The nanoscopic size of polymeric micelles<br />

makes the carrier unrecognizable <strong>by</strong> the phagocytic cells of the reticuloendothelial system (RES),<br />

elongating their blood circulation, and facilitating the carrier’s extravasation from tumor vasculature.<br />

The small size of polymeric micelles is also expected to ease penetration of the carrier within<br />

the tumor tissue and further internalization of polymeric micelles into the tumor cells. Hydrophobic<br />

core of polymeric micelles provides an excellent host for the incorporation and stabilization of anticancer<br />

agents that are mostly hydrophobic. Steric effect induced <strong>by</strong> the dense hydrophilic brush on<br />

the micellar surface protects the carrier against attachment of proteins on the micellar surface,<br />

avoiding early uptake and clearance of the carrier <strong>by</strong> RES leading to prolonged circulation time and<br />

higher accumulation of the carrier and the encapsulated drug in selective tissues that have leaky<br />

vasculature (e.g., tumor or inflammation sites). Polymeric micelles have been the focus of several<br />

reviews in recent years. 1–11<br />

Among different micelle-forming block copolymers developed to date, those with poly(ethylene<br />

oxide) (PEO) as the shell-forming block and poly(L-amino acid)s (PLAA)s and poly(ester)s as<br />

the core-forming block are in the front line of drug development. The placement is owed to the<br />

biocompatibility of the PEO and biodegradability of PLAA and poly(ester) structures. The primary<br />

advantage of PEO-b-PLAA block copolymers over PEO-b-poly(esters) is the chemical flexibility of<br />

the PLAA structure that makes nano-engineering of the carrier a feasible approach. 4 To date,<br />

research on PEO-b-PLAA for drug delivery has been mainly conducted on amino acids with<br />

functional side groups in their chemical structure, including L-aspartic acid, L-glutamic acid,<br />

L-lysine, and L-histidine. The presence of free functional side groups on the PLAA block provide<br />

sites for the attachment of drugs, drug compatible moieties, or charged therapeutics such as DNA.<br />

Moreover, a systemic alteration in the structure of the core-forming block also may be used to better<br />

control the extent of drug loading, release, or activation.<br />

Chemical modifications in the structure of the PLAA block have led to the development of<br />

optimal PEO-b-PLAA-based polymeric micellar formulation for the delivery of a number of potent<br />

therapeutic agents, such as doxorubicin (DOX), 12–23 paclitaxel (PTX), 24 cisplatin (CDDP), 25–32<br />

amphotericin B (AmB), 33–41 etc. To this point, three PEO-b-PLAA-based polymeric micellar<br />

formulations (all for the solubilization and delivery of anti-cancer agents) have successfully<br />

passed the phase of bench-top development and advanced to the stage of clinical evaluations<br />

(Table 18.1). 23,24,32 This chapter provides an overview on the preparation; micellar properties<br />

and biological performance of PEO-modified PLAA- based polymeric micellar formulations.<br />

Emphasis has been placed on the application of PEO-b-PLAA- micelles for the targeted delivery<br />

of anti-cancer agents. At the end, new advancements in the field, including a second generation of<br />

PEO-b-PLAA micelles (polymeric micelles for active targeting) and development of PEO-b-PLAA<br />

micelles for gene delivery are briefly discussed.<br />

18.2 SYNTHESIS OF PEO-b-PLAA BLOCK COPOLYMERS<br />

Nanotechnology for Cancer Therapy<br />

The traditional method for the synthesis of PLAA relies on the ring-opening polymerization (ROP)<br />

of a-amino acid-N-carboxyanhydrides (NCAs) (Figure 18.1). 42 NCAs can be pre-prepared from<br />

a-amino acids using a solution of phosgene in THF <strong>by</strong> the Fuchs–Farthing method. 43,44 NCA<br />

polymerization can be initiated with different nucleophiles or bases, the most common being<br />

primary amines and alkoxide anions. 45–48 This synthetic method is appealing to the pharmaceutical<br />

industry because it involves simple reagents, and it is economical; high molecular weight polymers<br />

can be prepared in both good yield and large quantity <strong>by</strong> this method.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 359<br />

TABLE 18.1<br />

Most Common PEO-Modified PLAA Micelles Investigated for Drug Delivery<br />

Drug Incorporation<br />

Method PLAA Block Incorporated Drug<br />

Latest Reported<br />

Phase of Progress References<br />

Chemical conjugation a<br />

P(Lys) Cyclophosphamid<br />

sulfide<br />

Development 96<br />

P(Asp) Doxorubicin Development 12, 18, 22, 98, 100<br />

PHEA Methotrexate Development 33, 87, 90<br />

Physical encapsulation b<br />

PBLA Doxorubicin Development 13, 94, 95<br />

P(Asp)-DOX Doxorubicin Clinical trials phase II 12, 17, 21, 23, 100<br />

PBLG Doxorubicin Development 133<br />

PPBA Paclitaxel Clinical trials phase I 24<br />

PBL-C16(Asp) KRN-5500 Development 65, 110<br />

PBLA Camptothecin Development 113<br />

P(n-butyl-L-Asp) Camptothecin Development 113<br />

P(lauryl-L-Asp) Camptothecin Development 113<br />

P(methylnaphtyl-L-<br />

Asp)<br />

Camptothecin Development 113<br />

PBLA Amphotericin B Development 40, 41<br />

PHSA Amphotericin B Development 33–39<br />

PBLA Indomethacin Development 64<br />

Electrostatic complexation c P(Asp) Cisplatin Development 27, 30, 31<br />

P(Glu) Cisplatin Clinical trials phase I 29, 32<br />

a<br />

Most common chemical structures shown in Figure 18.5.<br />

b<br />

Most common chemical structures shown in Figure 18.6.<br />

c<br />

Most common chemical structures shown in Figure 18.7.<br />

For the synthesis of PEO-b-PLAA block copolymers, a-methoxy-u-amino PEO has been used<br />

as the initiator. Synthesis of PEO-b-poly(b-benzyl-L-aspartate) (PEO-b-PBLA) and PEO-b-poly[3-<br />

(benzyloxycarbonyl)-L-lysine] (PEO-b-PBLL) from polymerization of b-benzyl-L-aspartate-Ncarboxyanhydride<br />

(BLA–NCA) and N-carboxyanhydride of 3-(benzyloxycarbonyl)-L-lysine<br />

(BLL–NCA), respectively, using a-methoxy-u-amino PEO as initiator has been reported. 49<br />

When primary amines are used as initiators, the NCA polymerization may proceed <strong>by</strong> two different<br />

mechanisms: amine mechanism and activated monomer (AM) mechanism (Figure 18.2A and<br />

Figure 18.2B). 48,50 The amine mechanism is a nucleophilic ring-opening chain-growth process<br />

where the polymer linearly grows with monomer conversion (Figure 18.2A). In the AM<br />

mechanism, NCA will be deprotonated, forming a nucleophile that initiates chain growth<br />

(Figure 18.2B). Polymerization <strong>by</strong> the AM mechanism may lead to side reactions. Besides, the<br />

initiator will not be part of the final product. In a given polymerization process, the system can<br />

R CH C<br />

NH 2<br />

O H N<br />

OH<br />

COCl 2<br />

O<br />

C<br />

O<br />

CH<br />

C<br />

R Nucleophile<br />

or base<br />

FIGURE 18.1 General scheme for the synthesis of PEO-b-PLAA-based block copolymers <strong>by</strong> ring-opening<br />

polymerization of a-amino acid-N-carboxyanhydrides (NCAs).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

H<br />

N<br />

R<br />

O<br />

n


360<br />

(a)<br />

(b)<br />

R' NH 2<br />

R' NH 2<br />

+<br />

R' NH3 N −<br />

R<br />

+<br />

O<br />

O<br />

O<br />

O<br />

R<br />

R'<br />

O<br />

N<br />

H<br />

N<br />

H<br />

O<br />

R H<br />

N<br />

R<br />

O O<br />

O<br />

O<br />

R<br />

Reaction with NCA<br />

N<br />

or N-aminoacyl NCA<br />

O<br />

O<br />

O<br />

O R<br />

N-aminoacyl NCA<br />

n NH2 NH 2<br />

R<br />

alternate between the amine and AM mechanisms. Because a propagation step for one mechanism<br />

is a side reaction for the other and vice versa, block copolymers prepared from the NCA method<br />

using amine initiators have structures different from those predicted <strong>by</strong> monomer feed compositions<br />

and, most likely, have considerable homopolymer contamination.<br />

To avoid side reactions, transition metal initiators have been developed 51–55 that use transition<br />

metal complexes as the end groups to control the addition of each NCA monomer to the polymer<br />

chain ends. However, the presence of chain-transfer reactions prevents the preparation of high<br />

molecular weight PLAAs <strong>by</strong> this process.<br />

Living polymerization is a relatively novel method to prepare PLAA that has been developed to<br />

overcome the mentioned limitations of existing methods (Figure 18.3). 56–58 In this method, the<br />

transition metal initiator activates the monomers and forms covalent active species that permit the<br />

formation of polypeptides via the living polymerization of NCAs. The metals react identically with<br />

NCA monomers to form metallacyclic complexes <strong>by</strong> oxidative addition across the anhydride bond<br />

of NCA. The AB diblock, PEO-b-poly(L-lysine) (PEO-b-P(L-Lys)), ABA triblock, poly(g-benzyl-<br />

L-glutamate)-b-PEO-b-poly(g-benzyl-L-glutamate) (PBLG-b-PEO-b-PBLG) copolymers, and<br />

diblock copolymers of poly(methyl acrylate)-b-(PBLG) of high molecular weights have been<br />

synthesized <strong>by</strong> living polymerization mechanism. 50,56,59<br />

Alternatively, PEO can be coupled to PLAA after the polymerization of PLAA. For instance,<br />

poly(L-histidine) (P(L-His)) has been synthesized <strong>by</strong> base-initiated ROP of protected NCA of L-His<br />

and coupled to carboxylated PEO to form PEO-b-P(L-His) via an amide linkage using dicyclohexyl<br />

carbodiimide (DCC) and N-hydroxysuccinimide (NHS). 60<br />

R '<br />

nCO 2<br />

H<br />

N<br />

R' NH 3 +<br />

N<br />

O<br />

nNCA<br />

O<br />

O<br />

O<br />

R<br />

R<br />

+ NCA<br />

O<br />

H O<br />

+ transfer<br />

−CO2 Further<br />

Oligopeptides<br />

O<br />

Condensation<br />

O<br />

Nanotechnology for Cancer Therapy<br />

N<br />

N<br />

H<br />

R'<br />

O<br />

N<br />

O<br />

O<br />

H<br />

N<br />

OH<br />

O<br />

R H<br />

N<br />

O O<br />

O<br />

O<br />

R<br />

H<br />

N<br />

O<br />

R H<br />

N<br />

R<br />

R<br />

C<br />

O<br />

n<br />

O<br />

NH 2<br />

O −<br />

R<br />

CO 2<br />

R' NH 3 +<br />

FIGURE 18.2 (a) Amine mechanism and (b) activated monomer (AM) mechanism of NCA polymerization.<br />

(From Deming, T. J., Advanced Drug Delivery Reviews, 54, 8, 2002. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

NH 2


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 361<br />

O<br />

2 NH<br />

(L)nM +<br />

3<br />

O 1<br />

4<br />

5<br />

M = Co, Ni<br />

R<br />

O<br />

NC A<br />

R<br />

R<br />

O<br />

−CO 2<br />

H<br />

N<br />

NCA<br />

R<br />

O<br />

M N H<br />

N<br />

H (L)n<br />

R<br />

HN<br />

M(L)n<br />

R<br />

n<br />

O<br />

R<br />

O<br />

N<br />

O<br />

N<br />

H<br />

O<br />

O<br />

N<br />

H<br />

R<br />

−nCO 2<br />

O-C 5<br />

HN<br />

(L) nM<br />

R<br />

(L)n M<br />

PEO-co-P(L-aspartic acid) (PEO-co-P(L-Asp))-bearing amine groups on the P(L-Asp) block<br />

have been synthesized <strong>by</strong> the melt polycondensation of N-(benzyloxycarbonyl)-L-aspartic acid<br />

anhydride (N-CBZ-L-ASP anhydride) and low molecular weight PEO. The product was an alternating<br />

copolymer having reactive amine groups on the P(L-Asp) residue. The backbone was linked<br />

<strong>by</strong> ester bond that is more biodegradable than the amide bond between PEO and P(L-Asp). 61<br />

Preparation of micelle-forming PEO-b-PLAA-based block copolymers with poly(leucine),<br />

poly(tyrosine), poly(phenylalanine) (PPA), and a composition peptide, i.e., poly(phenylalanineco-leucine-co-tyrosin-co-tryptophan)<br />

P(FLYW), as the core forming block through solid phase<br />

peptide synthesis has also been reported. 62 Methoxy PEO-b-PLAA dendrimers have also been<br />

prepared <strong>by</strong> the liquid phase peptide synthesis and used in drug delivery. 63<br />

18.3 MICELLIZATION OF PEO-b-PLAA BLOCK COPOLYMERS<br />

R<br />

NH<br />

proton migration<br />

O<br />

H<br />

N<br />

(L) nM<br />

FIGURE 18.3 Living polymerization of NCA.<br />

R<br />

O<br />

HN<br />

O O<br />

Assembly of micelle-forming PEO-b-PLAA block copolymers through one of the following<br />

methods has been reported.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

R<br />

O<br />

O<br />

O<br />

−CO 2<br />

R<br />

HN<br />

O<br />

O<br />

proton migration<br />

HN<br />

C<br />

M<br />

O<br />

NH<br />

R<br />

−CO 2<br />

R O<br />

N<br />

(L)n<br />

O<br />

N<br />

M<br />

(L)n<br />

R<br />

−CO<br />

H<br />

O<br />

R<br />

N<br />

R<br />

R<br />

NH<br />

NCA −CO 2<br />

O<br />

M<br />

O<br />

HN C<br />

R<br />

O<br />

(L) nM<br />

O<br />

R<br />

O<br />

N<br />

H<br />

O<br />

N<br />

H<br />

(L)n<br />

n<br />

NH<br />

N<br />

H<br />

O<br />

N<br />

O<br />

R<br />

R<br />

R<br />

O<br />

R<br />

N H<br />

R


362<br />

18.3.1 DIALYSIS METHOD<br />

In this method, block copolymer is dissolved in a water miscible organic solvent first. Then, this<br />

solution is dialyzed against water. The semi-permeable membrane keeps the micelles inside the<br />

dialysis bag, but it allows removal of organic solvent. Gradual replacement of the organic solvent<br />

with water, i.e., the non-solvent for the core-forming block, triggers self-association of block<br />

copolymers (Figure 18.4A). 13,35,40,64,65<br />

18.3.2 SOLVENT EVAPORATION METHOD<br />

In this process, the polymer is dissolved in a volatile organic solvent. The organic solvent is then<br />

removed completely <strong>by</strong> evaporation (usually under reduced pressure), leading to the formation of<br />

polymer films. Next, this film is reconstituted in an aqueous phase <strong>by</strong> vigorous shaking. 36,66 The<br />

solvent evaporation method of micellization cannot be utilized for block copolymers having large<br />

hydrophobic segments because the polymer film cannot be reconstituted easily in an aqueous phase<br />

for those structures (Figure 18.4B).<br />

18.3.3 CO-SOLVENT EVAPORATION METHOD<br />

In this approach, polymer is dissolved in a volatile water miscible organic solvent such as methanol.<br />

Then self-assembly is triggered <strong>by</strong> the gradual addition of aqueous phase (non-solvent for the coreforming<br />

block) to the organic phase. This step is followed <strong>by</strong> the evaporation of the organic<br />

co-solvent (Figure 18.4C). 33,34<br />

Polymer is dissolved<br />

in the organic solvent<br />

Gradual<br />

mixing with<br />

water<br />

(C)<br />

(A)<br />

(B)<br />

Gradual displacement of<br />

organic solvent with water<br />

through dialysis<br />

Solvent<br />

evaporation/film<br />

formation<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 18.4 Micellization of PEO-b-PLAA-based block copolymers through (A) dialysis, (B) solvent<br />

evaporation, and (C) co-solvent evaporation methods.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 363<br />

18.4 RATIONAL DESIGN AND FUNCTIONAL PROPERTIES OF PEO-b-PLAA<br />

MICELLES IN DRUG DELIVERY<br />

18.4.1 THE PEO SHELL<br />

PEO has a safe history of application as a pharmaceutical ingredient for the design of non-immunogenic<br />

peptides, carriers, or biomedical surfaces. 67,68 In PEO-b-PLAA, the presence of hydrophilic<br />

PEO chains introduces amphiphilicity, leading to the self-assembly of block copolymer and formation<br />

of association colloids. The PEO shell is expected to induce steric repulsive forces and to<br />

stabilize the colloidal interface against aggregation or absorption of proteins to the carrier as<br />

well. 10,69–71 As a result, the colloidal carrier will stay within the appropriate size range and<br />

surface properties that can avoid uptake <strong>by</strong> RES. A major barrier against targeted tumor delivery<br />

<strong>by</strong> colloidal carriers is the non-specific uptake and clearance of the carrier from the blood stream <strong>by</strong><br />

RES. Avoiding RES will prolong the circulation of polymeric micelles in blood, providing more<br />

chance for the extravasation of the carrier at sites with leaky vasculature (e.g., solid tumors).<br />

The extent of steric stabilization is found to be dependent on the length and density of the PEO<br />

chain on the surface of colloidal carriers. 20,72–74 Elongation of the PEO chain or an increase in the<br />

aggregation number of PEO-b-PLAA micelles may lead to the reduced chance of micellar aggregation<br />

and longer circulation time in blood and a higher probability of passive targeting to the tumor<br />

site for polymeric micelles. 10 The same parameter may reduce the adherence and interaction of the<br />

polymeric micelles with target cells. 75<br />

18.4.2 THE PLAA CORE<br />

Chemical flexibility of the PLAA structure as the core-forming block is the primary advantage of<br />

the PEO-b-PLAA-based polymeric micelles. Existence of several functional groups on a PLAA<br />

block provides a number of sites for the attachment of drugs, drug compatible moieties, or charged<br />

therapeutics to a single polymeric backbone in micelle-forming PEO-b-PLAA block copolymers.<br />

This may lead to a lower dose of administration for drugs that are chemically conjugated to the<br />

PLAA block of PEO-b-PLAA. On the other hand, the diversity of functional groups in a PLAA<br />

chain (amino, hydroxyl, and carboxylic groups) allows conjugation of different chemical entities to<br />

the polymeric backbone and provides an opportunity for the design of polyion complex and/or pH<br />

responsive polymeric micelles. Finally, through changes in the chemical structure of the PLAA<br />

block, it will also be possible to fine tune the structure of PEO-b-PLAA micelles to achieve optimal<br />

properties for drug targeting. 4 From a pharmaceutical point of view, PEO-b-PLAA micelles are<br />

considered advantageous over PEO-b-poly(ester)s because they can easily be reconstituted after<br />

freeze–drying and do not need lyoprotection.<br />

The most effort for the production of PEO-b-PLAA micellar delivery systems has been focused<br />

on the application of P(L-Asp), poly(L-glutamic acid) (P(L-Glu)), and P(L-Lys)-based polymers<br />

(Table 18.1, Figure 18.5 through 18.7). This is owed to the presence of free carboxyl and amino<br />

groups that makes these structures useful for the chemical conjugation and electrostatic complexation<br />

of drugs and DNA to the PLAA block. 76<br />

A major concern for the application of PLAA-based pharmaceuticals is the possibility of immunogenic<br />

reactions and long-term accumulation and toxicity <strong>by</strong> these agents. Unfortunately, the<br />

information on the biocompatibility and biodegradability of PLAAs is limited. 77,78 Nevertheless,<br />

the results of ongoing clinical trials in Japan on PEO–PLAA-based micelles for the delivery of DOX,<br />

PTX, and CDDP is expected to provide more information on the safety of these structures.<br />

18.4.3 MICELLAR DIMENSIONS<br />

PEO-b-PLAA micelles developed to date are mostly in a diameter range of 10–100 nm. 79 At this<br />

specific size range, PEO-b-PLAA micelles are expected to be big enough to avoid glomerular<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


364<br />

O<br />

O<br />

PEO-b-P(Asp)-DOX CH3 O CH2 CH2 NH<br />

n<br />

C CH<br />

CH2 NH<br />

x<br />

C CH2 CH NH H<br />

y<br />

C O<br />

C O<br />

R<br />

R = OH or<br />

O<br />

filtration and elimination <strong>by</strong> kidney and sufficiently small to escape RES clearance. 80,81 As a result,<br />

polymeric micelles may stay in circulation in the blood for prolonged periods, decreasing the<br />

clearance (CL), and volume of distribution (Vd), but increasing the area under the plasma concentration<br />

versus time curve (AUC) of the encapsulated drug if the drug stays within the carrier in<br />

biological system. This change in the pharmacokinetic profile may allow for a reduction in the dose<br />

of administration and/or decreased toxicity for the incorporated drug.<br />

Carriers of less than 200 nm with prolonged blood circulation properties are shown to passively<br />

accumulate in solid tumors. 82 The new blood vessels in the tumor, formed to provide access to<br />

nutrients and oxygen, are usually poorly made and have large gaps (150–730 nm) in their endothelium.<br />

This allows the extravasation of nanoparticles of the appropriate size range to the<br />

extravascular space surrounding the tumor cells. 83,84 Because the lymphatic system that drains<br />

fluids out of other organs is absent in tumors, the permeated nanocarrier is expected to become<br />

trapped in the tumor. This phenomenon, known as the enhanced permeation and retention (EPR)<br />

effect, is believed to be the reason for the passive accumulation of colloidal carriers (e.g., liposomes,<br />

polymerosomes, and polymeric micelles) in solid tumors. 85,86 The polymeric micelles’<br />

smaller size compared to other colloidal carriers is expected to ease further penetration of the<br />

delivery system inside the dense environment of solid tumors. Other advantages associated with<br />

the small dimensions of polymeric micelles include the ease of sterilization via filtration and safety<br />

of administration.<br />

R<br />

OH<br />

OH O OH<br />

O<br />

CH3 O<br />

OH<br />

NH<br />

PEO-b-PHEA-MTX CH3 O CH2 CH2 CH2 m<br />

CH2 NH C CH NH<br />

CH2 C CH NH<br />

x<br />

CH2 H<br />

y<br />

HOOC CH2 CH2 CO O CH2 CH2 NH CO<br />

CO NH CH2 CH2 OH<br />

H<br />

NH<br />

C<br />

O<br />

N CH 3<br />

CH 2<br />

FIGURE 18.5 Chemical structure of some micelle-forming PEO-b-PLAA drug conjugates designed for drug<br />

delivery.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

N<br />

N<br />

O<br />

N<br />

OH<br />

OH<br />

NH 2<br />

Nanotechnology for Cancer Therapy<br />

N<br />

O<br />

NH 2


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 365<br />

O O<br />

PEO-b -P(L-Asp)-DOX CH3 OCH2CH2 NH<br />

n<br />

2 C CH NH C<br />

x<br />

CH2 CH2 CH NH H<br />

y<br />

C O<br />

C<br />

R<br />

O<br />

R<br />

O OH O<br />

OH<br />

PEO-b -PBLA<br />

PEO-b -P(n-butyI-L-Asp)<br />

PEO-b -P(lauryI-L-Asp)<br />

CH 3<br />

R =<br />

R =<br />

PEO-b -P(methynaphtyI-L-Asp)<br />

PEO-b -P(BLA, C16)<br />

R=OH or<br />

R =<br />

OH O OH<br />

O<br />

O<br />

CH3 OH<br />

NH<br />

O<br />

O CH2 CH2 NH<br />

n<br />

C CH<br />

CH2 NH H<br />

m<br />

CO O CH2 O<br />

O<br />

CH3 O CH2 CH2 CH2 n<br />

NH C CH NH C<br />

x<br />

CH2 CH NH H<br />

y<br />

CH2 C O R C O R<br />

CH2 CH3 3<br />

or H<br />

O<br />

O<br />

O<br />

O<br />

CH3 O CH2 CH2 CH<br />

n<br />

2 NH C CH NH<br />

x<br />

C CH2 CH NH H<br />

y<br />

CH2 CH3 11<br />

or H<br />

O<br />

CH C O R<br />

2<br />

O<br />

O<br />

C<br />

O<br />

O R<br />

CH3 O CH2 CH2 CH<br />

n<br />

2 NH C CH NH<br />

x<br />

C CH2 CH NH H<br />

y<br />

H2C or H<br />

CH2 CO R<br />

O<br />

C<br />

O<br />

O R<br />

CH3 O CH NH NH<br />

2 CH2 C CH C CH NH H<br />

n x m-x<br />

CH2 CH O C<br />

3 CH2 CH2 CH2 15<br />

CO O CH2 O<br />

FIGURE 18.6. Chemical structure of most commonly used PEO-b-PLAA-based polymers for physical encapsulation<br />

of drugs.<br />

The downside of polymeric micelles’ smaller size is the restricted space for drug encapsulation<br />

in the carrier. Second, this provides a limitation in the polymeric micelles’ ability for a sustained<br />

drug release because of a large total surface area, and finally, it allows for possible carrier accumulation<br />

in normal tissues bearing leaky vasculature such as the liver. Chemical engineering of the<br />

core structure in PEO-b-PLAA micelles may be used to modify the final dimensions of the carrier to<br />

an optimum scale so the polymeric micellar carrier can avoid each of these problems. Elongation of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

OH<br />

O


366<br />

PEO-b -PHSA<br />

PEO-b -PBLG<br />

PEO-b -PBLG star block<br />

copolymer<br />

PEO-b -PPA<br />

CH3 O CH2 CH2 O CH2 S CH<br />

n 2<br />

3<br />

H3C CH2 C O<br />

NH<br />

CH2 C<br />

NH<br />

CH<br />

CH2 C<br />

NH<br />

y<br />

C CH NH H<br />

z<br />

CH2 C NH CH2OH 6<br />

16<br />

O<br />

6<br />

O<br />

O<br />

C<br />

H 3 C CH 3<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

CH 3<br />

CH 2<br />

CH 2<br />

NH<br />

COO<br />

O<br />

O<br />

NH C CH NH<br />

CH2 CH2 O CH2 CH2 NH<br />

n<br />

w H<br />

CH<br />

2 2<br />

C O CH<br />

2<br />

C H<br />

6 5<br />

O<br />

CH2<br />

O<br />

O<br />

C CH NH<br />

CH 2<br />

CH 2<br />

C<br />

O<br />

H<br />

m<br />

O CH2 CH2 CH O CH2 CH2 O CH2 CH2 C O NH CH2 CH2 NH C CH NH H<br />

n k<br />

O<br />

O<br />

O<br />

C O CH 2 C 6H 5<br />

CH2 CH2 O CH2 CH2 CH O CH2CH2 O CH2 m<br />

COO<br />

O<br />

NH<br />

CH2 C O NH CH2 CH2 O<br />

NH C CH NH H<br />

s<br />

CH2 CH2 CH2 C O CH2C6H5 C O CH2 C6H5 CH2 CH2 2<br />

O<br />

NH C CH NH<br />

O<br />

H<br />

y<br />

CH3 O CH2 CH2 NH<br />

n<br />

O<br />

C<br />

CH<br />

NH<br />

Figure 18.6 (continued)<br />

Nanotechnology for Cancer Therapy<br />

H<br />

m<br />

O


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 367<br />

O<br />

O<br />

PEO-b -P(L-Asp)-CDD P CH3 O CH2 CH2 n<br />

NH C CH<br />

CH2 NH<br />

x<br />

C CH2 CH NH<br />

y<br />

C O<br />

H<br />

R O C<br />

O<br />

PEO-b -P(L-Glu)-CDD P<br />

the PLAA block and an increase in the level of substitution or the length of the side chain have<br />

shown an increasing effect on the final size of the prepared polymeric micelle. 34,37,87<br />

18.4.4 MICELLAR STABILITY<br />

CH3 O CH2 CH2 NH C CH NH H<br />

n x<br />

CH2 R = Na + or<br />

FIGURE 18.7 PEO-b-PLAA-based polymers used for the formation of polyion complex micelles.<br />

Thermodynamically and kinetically stable polymeric micellar structures may retain their integrity<br />

in a biological environment for longer periods and, more effectively, avoid uptake <strong>by</strong> RES and<br />

elimination through the kidney and possibly change the normal organ distribution of the encapsulated<br />

drug the same way. PEO-b-PLAA block copolymers are reported to be thermodynamically<br />

stable and characterized <strong>by</strong> very low critical micellar concentrations (CMC) in mM range. 37,87,88<br />

Therefore, even after intravenous application and dilution in blood, PEO-b-PLAA micelles may<br />

still stay above CMC and in a micellar form. The chemical structure of the PLAA core may be<br />

modified toward more hydrophobicity to push the CMC to even lower values. 35,37<br />

Formation of hydrophobic, electrostatic, and hydrogen bonds in the core of PEO-b-PLAA<br />

micelles after assembly of block copolymers makes the micellar structure kinetically stable. As<br />

a result, polymeric micelles may not necessarily exist in equilibrium with polymeric unimers above<br />

CMC, and their dissociation will be very slow below CMC. The degree of kinetic stability is<br />

suggested to be high, especially for block copolymer structure with stiff or bulky core-forming<br />

blocks. Cross-linking of the core structure may enhance the kinetic stability of polymeric micelles<br />

as well. 89<br />

Currently, there has been no report on the development of experimental tools that can assess<br />

directly the in vivo stability of micellar structures. Instead, micellar stability is assessed indirectly<br />

through in vitro drug release and micellar dissociation studies under diluting condition of gel<br />

permeation chromatography (GPC). For instance, the stability of methotrexate (MTX) esters of<br />

PEO-block-poly(2-hydroxyethyl-L-aspartamide) (PEO-b-PHEA) (Figure 18.5) was investigated <strong>by</strong><br />

GPC, in vitro. At high drug conjugation levels, MTX conjugate micelles were shown to be quite<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

R<br />

H 3 N<br />

H 3 N<br />

O<br />

Pt<br />

C<br />

O<br />

O<br />

O<br />

Cl<br />

CH 2<br />

or<br />

H 3 N<br />

H 3 N<br />

Pt<br />

R


368<br />

stable and entirely elute as micelles during GPC. Likewise, the level of stearic acid substitution in<br />

micelles of PEO-block-poly(N-hexyl stearate-L-aspartamide) (PEO-b-PHSA) (Figure 18.6) was<br />

shown to influence micellar stability during in vitro studies. 87 A comparison between the pharmacokinetics<br />

and biodistribution of polymeric micellar carrier-encapsulated drug and free drug (or<br />

commercially available solubilized form of the drug) in animal models has provided an indirect<br />

measure for the evaluation of the carrier’s in vivo stability. For example, the presence of DOX<br />

dimers in PEO-b-P(Asp)-based formulation of DOX has been found to improve the pharmacokinetic<br />

and biodistribution profile of DOX for passive targeting. This effect has been attributed to the<br />

stabilization of the micellar formulation <strong>by</strong> DOX dimmers. 17 A similar effect is observed for CDDP<br />

polymeric micellar formulations when the structure of the core-forming block is made more<br />

hydrophobic <strong>by</strong> changing PLAA structure from P(L-Asp) to P(L-Glu) (Figure 18.7). 29<br />

18.4.5 DRUG INCORPORATION AND RELEASE PROPERTIES<br />

Nanotechnology for Cancer Therapy<br />

Incorporation of drugs <strong>by</strong> PEO-b-PLAA micelles has been accomplished through three different<br />

means: chemical conjugation of drug to the PLAA block and formation of micelle-forming polymer/drug<br />

conjugates; physical encapsulation of drugs in PEO-b-PLAA micelles; and formation of<br />

an electrostatic complex between the PLAA core and charged drug followed <strong>by</strong> the self-assembly<br />

of the polyion complex (Table 18.1, Figure 18.5 through 18.7).<br />

Drug cleavage from the polymeric backbone <strong>by</strong> non-specific hydrolysis or enzymes is a prerequisite<br />

for the activity of drug-block copolymer conjugates. If the micellar structure can withstand<br />

dissociation in a biological environment, then the hydrophobic core of polymeric micelles may<br />

protect the degradable bonds between the polymeric backbone and drug from exposure to water and<br />

hydrolysis. In this case, the micelle-forming block copolymer drug conjugate may prevent the<br />

premature release of the incorporated drug in blood circulation and provide a delayed mode of<br />

release for the cytotoxic agent that is only triggered after accumulation and uptake of the carrier <strong>by</strong><br />

tumor cells. The other possible scenario will be the excessive stability of the polymeric pro-drug<br />

that may lead to the inactivity of the final product. Incorporation of hydrophilic moieties, formation<br />

of pH sensitive bonds between polymer and drug, and changes in the molecular weight of the coreforming<br />

block may be used to modify micellar stability and drug release properties of micelleforming<br />

PEO-b-PLAA-based drug conjugates. 87,90,91<br />

Generally, physical encapsulation of drugs within polymeric micelles is a more attractive<br />

approach because many drug molecules do not bear reactive functional groups for chemical conjugation.<br />

In this system, formation of hydrophobic interactions or hydrogen bonds between the<br />

micelle-forming block copolymer and physically encapsulated drug provides basis for drug solubilization<br />

and stabilization within the micellar carrier. Unlike polymer–drug conjugates that may<br />

provide a delayed release for the incorporated drug, physical encapsulation of hydrophobic drugs in<br />

polymeric micelles usually leads to a sustained mode of drug release from the carrier. It is possible<br />

that the drug content of physically loaded polymeric micelles is prematurely released in the blood<br />

circulation, i.e., before the carrier reaches its target. In this case, the released drug and carrier will<br />

follow separate paths in the biological system, and the colloidal carrier will be ineffective in terms<br />

of drug targeting. To lower the rate of micellar dissociation, drug diffusion, and the overall rate of<br />

drug release from the micellar carrier, the PLAA core of PEO-b-PLAA micelles may be tailored<br />

either to bear more drug-compatible moieties, to become glassy under physiological condition<br />

(378C), or to become cross-linked. 16,33,38,39,65,89,92–94 Finally, the method of drug incorporation<br />

in polymeric micelles also may be modified to improve the extent of drug loading, localization, or<br />

physical state of the loaded drug providing other means for controlling the rate of drug release from<br />

polymeric micelles. 36,66,95<br />

Polyion complex micelles can incorporate charged therapeutics through electrostatic<br />

interactions between oppositely charged polymer/drug combinations. Neutralization of the<br />

charge on the core-forming segment of the block copolymer will trigger self-assembly of the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 369<br />

polyion complex and further stabilization of the complex within the hydrophobic environment of<br />

the micellar core. The extent of drug release from polyion complex micelles is dependent on the<br />

rate of drug exchange with free ions and proteins in the physiological media. The presence of ions<br />

and proteins at high concentrations in the biological environment has been shown to lead to a<br />

relatively fast rate of drug release and micellar instability for polyion complex micelles in vivo. In<br />

this context, chemical manipulation of the PLAA core may be used to increase the hydrophobicity<br />

and rigidity of the micellar core, restricting the penetration of free ions to the micellar core in<br />

polyion complex micelles. 26,29 In addition, the micellar core may be cross-linked to avoid its<br />

dissociation. 89,92,93 This may lead to a sustained or even delayed mode of drug release from<br />

the carrier.<br />

18.5 PEO-b-PLAA MICELLES FOR THE DELIVERY OF ANTI-CANCER DRUGS<br />

18.5.1 CYCLOPHOSPHAMIDE<br />

The first attempt for the design of PEO-b-PLAA-based micellar systems for targeted delivery of<br />

anti-cancer drugs was made <strong>by</strong> Ringsdorf et al. in the 1980s. 96 In this study, cyclophosphamide<br />

(CP) sulfide and palmitic acid were attached to the P(L-Lys) block of a PEO-b-P(L-Lys). Palmitic<br />

acid was used as a hydrophobic moiety to induce the required amphiphilicity for micelle-formation<br />

to the block copolymer. CP is an alkylating agent that is transformed to active alkylating metabolites<br />

via hepatic and intracellular enzymes. The polymeric micellar formulation was found to be<br />

efficient in the stabilization of active CP metabolite, and it caused a five-fold increase in the life<br />

span of L1210 tumor-bearing mice even at a reduced CP-equivalent dose.<br />

18.5.2 DOXORUBICIN<br />

The PEO-b-PLAA-based micellar formulation of DOX was developed <strong>by</strong> Kataoka et al. in the late<br />

1980s and early 1990s and then entered clinical trials in 2001 in Japan. In early studies, micelleforming<br />

PEO-b-P(L-Asp) conjugates of DOX (Figure 18.5) were synthesized and tried in different<br />

in vitro and in vivo cancer models for their pharmacokinetics, biodistribution, toxicity, and efficacy.<br />

18,22,97–100 The results of these studies pointed to a higher blood and tumor levels for PEO-b-<br />

P(L-Asp)-DOX micelles compared to free drug, especially for micelles having longer PEO<br />

chain lengths. A tumor/heart concentration ratio of 12, 8.1, and 0.9 was demonstrated for PEOb-P(L-Asp)-DOX<br />

with PEO chains of 12000, 5000 gmol K1 , and free DOX, respectively, 24 h after<br />

intravenous injection. Cardiomyopathy is the most important and dose-limiting adverse effect of<br />

DOX. The maximum tolerable dose (MTD) of DOX was increased (almost 20-fold) <strong>by</strong> its PEO-b-<br />

P(L-Asp) conjugate that allowed administration of higher drug levels. However, compared to free<br />

drug, higher levels of conjugated DOX were required for an equal anti-tumor activity in vitro and<br />

in vivo. The formulation was found to contain some un-conjugated, physically entrapped DOX that<br />

was, in fact, responsible for the anti-tumor activity of this system.<br />

In further studies, PEO-b-P(L-Asp)-DOX micelles were utilized as micellar nanocontainers for<br />

the physical encapsulation of DOX. 12,16,17 To delay the release of DOX unimers from the formulation<br />

in a biological environment, DOX dimers were co-incorporated with DOX unimer within the<br />

micellar core. Compared to free DOX, within 24 h, DOX encapsulated in PEO-b-P(L-Asp)-DOX<br />

micelles showed 28.9-fold higher AUC in plasma, 3.4-fold higher tumor AUC, less toxicity, and<br />

superior in vivo activity in solid and hematological cancers in mice. After 24 h, the pharmacokinetics<br />

of encapsulated and free DOX became similar, pointing to drug release from the<br />

micellar carrier.<br />

Because lyophilized micelles that contained DOX dimers were not reconstitutable in water after<br />

long periods of storage, the dimer form of DOX has been removed from this formulation for clinical<br />

trials. 21,23 The physically encapsulated DOX in PEO-b-P(L-Asp)-DOX micelles, i.e., NK911,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


370<br />

showed similar side effects to free DOX in phase I clinical trails. In pharmacokinetic evaluations in<br />

humans, NK911 exhibited 2.5-fold increase in half-life, 2.2-fold decrease in CL, 1.6-fold decrease<br />

in Vd, and 2-fold increase in plasma AUC compared to free DOX. The lower degree of change in<br />

pharmacokinetic parameters for the NK911 compared to liposomal formulation of DOX pointed to<br />

the lower stability of polymeric micellar formulation. However, infusion-related symptoms<br />

observed after administration of liposomal DOX were not seen for NK911. This formulation has<br />

entered phase II clinical trials for the treatment of metastatic pancreatic cancer.<br />

Another PEO-b-PLAA-based micellar structure with benzyloxy group as the core-forming<br />

block has also been used for physical encapsulation of DOX <strong>by</strong> the same research group. 13,94,95<br />

PEO-b-PBLA (Figure 18.6) efficiently encapsulated DOX and sustained its rate of release (50%<br />

drug release within 72 h). The formulation was not as effective as PEO-b-P(L-Asp)-DOX carrier of<br />

DOX in animal models in reducing the toxicity of DOX. MTD of DOX was raised 2.3-fold <strong>by</strong> PEOb-PBLA<br />

in C-26-bearing mice (compared to a 20-fold increase in MTD observed for the PEO-b-<br />

P(L-Asp)-DOX carrier).<br />

18.5.3 PACLITAXEL<br />

PTX is a poorly water-soluble drug (water solubility of 0.4 mg/mL). For clinical use, PTX is<br />

solubilized using a significant amount of Cremophor EL in Taxol w formulation (Bristol–Myers<br />

Squibb). Several attempts have been made to develop safe solubilizing agents for this important<br />

chemotherapeutic agent. 1 However, the problems of low solubility, drug leakage, and precipitation<br />

upon dilution in biological system for PTX have not been completely resolved.<br />

The hydrophobic nature of PTX and the absence of safe carriers that can reduce the toxicity and<br />

enhance the efficacy of injectable PTX have driven a tremendous amount of attention to the<br />

application of polymeric micellar carriers for this drug.<br />

The only polymeric micellar formulation showing benefit in drug targeting for PTX is its PEOb-PLAA-based<br />

micellar formulation, namely NK105. Recently, Hamaguchi et al. reported on the<br />

development of a PEO-b-PLAA-based polymeric micellar formulation for PTX delivery. 24 The<br />

formulation consists of physically encapsulated PTX in micelles of PEO-b-poly(4-phenyl-1-butanoate)L-aspartamide<br />

(PEO-b-PPBA). In preclinical studies in animal models, NK105 has shown<br />

86-fold increase in its AUC in plasma, 86-fold decrease in CL, and 15-fold decrease in steady-state<br />

volume of distribution (Vdss) compared to Taxol w after intravenous injection. 24 This has resulted<br />

in a 25-fold increase in drug AUC in tumor and stronger anti-tumor activity in C-26 tumor-bearing<br />

mice models. At a PTX equivalent dose of 100 mg/kg, a single administration of NK105 resulted in<br />

the disappearance of tumors, and all mice remained tumor-free thereafter. This formulation is<br />

currently in phase I clinical trials in Japan.<br />

PTX polymeric micellar formulations based on PEO-b-poly(ester) block copolymers have also<br />

entered clinical evaluations in Canada and South Korea. 101 Although PEO-b-poly(ester) micelles<br />

were remarkably successful in raising the solubility of PTX, they have mostly failed in retaining<br />

their drug content in biological environment, there<strong>by</strong> regarded as ineffective in terms of drug<br />

targeting. 102–107<br />

18.5.4 CISPLATIN AND ITS DERIVATIVE<br />

Nanotechnology for Cancer Therapy<br />

CDDP is another anti-cancer agent that has been tried for encapsulation in polymeric micelles.<br />

Polyion complexes of positively charged CDDP and negatively charged free carboxyl group of the<br />

PEO-b-P(L-Asp) block copolymer were formed and assembled to micellar structures <strong>by</strong> Kataoka et<br />

al. (Figure 18.7). 26–31,76 Incorporation of CDDP in polymeric micelles has led to modest changes in<br />

the pharmacokinetic and biodistribution properties of CDDP in Lewis Lung carcinoma mice.<br />

Compared to free drug, 5.2- and 4.6-fold increases in the plasma and tumor AUC (calculated<br />

based on total platinum content) were observed for CDDP-loaded micelles. The low degree<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 371<br />

of change in pharmacokinetics of CDDP was attributed to the instability of the micellar<br />

formulation.<br />

In further studies, to restrict the ion exchange rate with the micellar core and stabilize the<br />

polymeric micellar formulation, the P(L-Asp) core of PEO-b-PLAA micelles was replaced with the<br />

more hydrophobic structure of P(L-Glu). 29 This change resulted in an increase in platinum levels in<br />

plasma and tumor for CDDP. In anti-tumor activity studies, four out of ten C-26-bearing mice<br />

receiving the PEO-b-P(L-Glu) formulation of CDDP showed complete cure, whereas, with free<br />

drug, no mice showed complete tumor regression. Recently, this formulation (named NC-6004) has<br />

concluded the stage of preclinical assessments showing 65-fold increase in plasma AUC, 9-fold<br />

decrease in total CL, and decreased nephro and neurotoxicity in rats. NC-6004 was also found to be<br />

equally effective as free CDDP against MKN-45 (human gastric cancer) tumor xenografts in mice.<br />

This formulation has recently moved to clinical trial in Japan. 32<br />

PEO-b-P(L-Glu) micelles have also been tried as carriers for a less toxic but more hydrophobic<br />

analog of CDDP, i.e., dichloro(1,2-diaminocyclohexane)platinum(II) (DACHPt) with hopes to<br />

increase the stability of polyion complex micellar formulation. In pharmacokinetic studies, a<br />

slightly higher platinum plasma levels for this system was observed compared to CDDP formulation,<br />

but tumor accumulation was similar for both drugs. 108<br />

18.5.5 METHOTREXATE<br />

Kwon et al. first reported on the conjugation of MTX to a PEO-b-P(L-Asp) derivative in 1999. 90 In<br />

this study, MTX was attached through an ester bond to PEO-b-PHEA, obtained <strong>by</strong> aminolysis of<br />

PEO-b-PBLA with ethanolamine. The yield of the reaction was greater than 90%, and MTX content<br />

was 9–20 wt%. The PEO-b-PHEA–MTX conjugates (Figure 18.5) showed self-assembly in an<br />

aqueous phase using a dialysis method that resulted in micelles with an average size of 14 nm<br />

and a narrow size distribution. For the in vitro release, a sample in a dialysis bag was placed into<br />

0.1 M phosphate buffer at varied pH (2.0–9.9). Free MTX was quickly released from the dialysis<br />

bag within 10 h. The in vitro release of MTX from self-assembled PEO-b-PHEA–MTX conjugate<br />

was slowest at pH 5 and somewhat faster at pH 2.2 and 7 (less than 20% of the drug was released<br />

over 260 h). At pH 10, release of MTX was rapid and almost complete in 80 h.<br />

In 2000, the same authors continued their study on this micellar formulation <strong>by</strong> changing the<br />

level of MTX substitution on the polymeric chain. 87 They obtained three different MTX substitution<br />

levels of 3.2, 9.0, and 19.4 wt% and showed a decrease in the CMC of PEO-b-<br />

PHEA–MTX with an increase in MTX substitution degree. The average diameter of micelles<br />

also increased from 11.7 to 27.1 nm as the level of MTX substitution was raised from 3.2 to<br />

19.4 wt%. The results of in vitro release experiments in PBS pH 7.4 showed a slower release<br />

rate at higher MTX substitution levels. For PEO-b-PHEA–MTX with 3.2, 9.0, and 19.4% of<br />

MTX substitution, 21, 10, and 5% of drug was released after 20 days, respectively.<br />

18.5.6 KRN5500<br />

KRN5500 is a new anti-cancer drug currently in clinical trials in Japan and the USA. It is a semisynthetic,<br />

water-insoluble analog of spicamycin derived from Streptomyces alanosinicus.<br />

KRN5500 is an inactive prodrug that is metabolized to 4-N-glycylspicamycin amino nucleoside<br />

(SAN-Gly) <strong>by</strong> a cytosomal enzyme. Fatty acid chains present in the chemical structure of<br />

KRN5500 are lost in SAN-Gly. As a result, SAN-Gly does not cross cellular membranes as<br />

easily as KRN5500 and is found to be 1000-fold less cytotoxic than KRN5500 in vitro. 109<br />

Development of a polymeric micellar formulation for KRN5500 was reported <strong>by</strong> Yokoyama<br />

et al. 65,110 To encapsulate KRN5500 that contains an aliphatic residue in its chemical structure, the<br />

aromatic group of the core-forming block in PEO-b-PBLA has been replaced with cetyl esters<br />

(Figure 18.6). Polymeric micellar KRN5500 and free drug were found to be similar in terms of antitumor<br />

activity against HT-29 (human colonic cancer) and MKN-45 xenografts in a mouse model,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


372<br />

but polymeric micellar formulation was less toxic. The vascular damage with fibrin clot or an<br />

increase in the plasma levels of BUN seen after intravenous administration of free drug was not<br />

observed for the polymeric micellar formulation. In a bleomycin-induced lung injury rat model,<br />

free KRN5500 at a dose of 3 mg/kg caused extensive pulmonary pathological changes with widespread<br />

hemorrhage; however, after administration of polymeric micellar formulation at a similar<br />

dose, no pathological change was observed, and the lung resembled the control group.<br />

18.5.7 CAMPTOTHECIN<br />

Camptothecin (CPT) is a naturally occurring cytotoxic alkaloid isolated from Camptotheca accuminata.<br />

It exists in two forms: biologically active lactone form and inactive carboxylate form. 111,112<br />

The lactone form of CPT exhibits poor aqueous solubility; therefore, it is a good drug candidate for<br />

incorporation in polymeric micelles.<br />

Camptothecin has been physically encapsulated in micelles of PEO-b-P(L-Asp) having benzyl,<br />

n-butyl, lauryl, and methylnaphtyl attached to the core-forming block through dialysis, emulsion,<br />

and solvent evaporation methods. 113 The presence of aromatic structures, e.g., benzyl or methylnaphtyl<br />

groups, and application of a solvent evaporation method of physical encapsulation were<br />

shown to be more efficient, resulting in the stabilization of encapsulated CPT in the micellar<br />

structures. The optimized polymeric micellar formulation of CPT was shown to stabilize the<br />

lactone form of CPT even in the presence of a stimulated physiological condition. In the presence<br />

of serum, 72% of CPT structure was protected <strong>by</strong> polymeric micelles, but free drug lost 80% of its<br />

active form.<br />

18.6 PEO-b-PLAA MICELLES FOR THE DELIVERY OF OTHER THERPEUTIC<br />

AGENTS<br />

18.6.1 AMPHOTERICIN B<br />

Nanotechnology for Cancer Therapy<br />

AmB is the most potent anti-fungal agent used in systemic mycosis. It is an amphiphilic drug with<br />

very low water solubility (aqueous solubility of 1 mg/mL). For clinical use, it is solubilized with the<br />

aid of a low molecular weight surfactant, sodium deoxycholate, in Fungizone w . It is also available<br />

as three lipid-based formulations: a lipid complex (Ablect w ), a complex with cholesteryl sulfate<br />

(Amphotec w ), and a conventional liposome (Ambisome w ).<br />

Micelles of PEO-b-PHSA (Figure 18.6) having aliphatic structures in their core were prepared<br />

and used for the solubilization of AmB through a solvent evaporation method. 33–36,38,39 In this<br />

system, the level of aliphatic chain substitution was found to affect the encapsulation efficiency and<br />

release rate of AmB where higher levels of aliphatic side chains in the micellar core led to higher<br />

AmB encapsulation and a lower rate of drug release. As a result of a sustained mode of drug release,<br />

hemolytic activity of AmB was reduced in its polymeric micellar formulation in comparison to free<br />

AmB or Fungizone w . The anti-fungal efficacy of polymeric micellar formulation was found to be<br />

comparable to that of free AmB. In further studies, the hemolytic activity of AmB incorporated in<br />

PEO-b-PLAA micelles having different fatty acid chain lengths as the core-forming block (at a<br />

range between 2 and 18 carbons) but similar substitution levels (86–95%) were compared. The<br />

results pointed to an advantage for stearic acid (saturated fatty acid with an 18 carbon chain length)<br />

in terms of reduced AmB hemolytic activity. In vivo efficacy studies of this formulation in a<br />

neutropenic murine model of disseminated candidiasis indicated a similar efficacy for PEO-b-<br />

PHSA formulation of AmB and Fingizone w . The physical encapsulation of AmB in PEO-b-<br />

PBLA at an alkaline pH has also been reported. 40,41<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 373<br />

18.6.2 INDOMETHACIN<br />

Indomethacin is a non-steroidal anti-inflammatory drug (NSAID), showing a low solubility in water<br />

(35 mg/mL). A properly designed polymeric micellar carrier may increase the solubility of indomethacin,<br />

control its renal side effects, and increase accumulation of this drug in inflamed areas.<br />

PEO-b-PBLA micelles have been used to solubilize indomethacin, and they showed a sustained<br />

mode of release for the encapsulated drug. The release rate of drug from PEO-b-PBLA micelles was<br />

found to be dependent on the ionization state of indomethacin where maximum control on drug<br />

release was observed at pH values below the pK a of indomethacin (4.5). At this condition, indomethacin<br />

was unionized and favored the non-polar environment of PBLA core in polymeric<br />

micelles. 64<br />

18.7 PEO-b-PLAA MICELLES FOR ACTIVE DRUG TARGETING<br />

The clinical advantage of polymeric micelles in terms of drug targeting may be further improved<br />

if the carrier can take advantage of differences between cancer and normal tissue at a cellular level.<br />

This may be achieved through the attachment of targeting ligands that recognize cancerous tissue<br />

to the surface of polymeric micelles, rendering them effective in active drug targeting. PEO-b-<br />

PLAA micelles are excellent candidates for active drug targeting because chemical modifications<br />

on both core- and shell-forming blocks are possible in PEO-b-PLAA-based carriers. Targeting<br />

ligands can be attached at the end of the PEO chains, and the anti-cancer drugs can be physically<br />

or chemically entrapped in the core. To date, various ligands such as different sugars, transferrin,<br />

folate residues, and peptides have been attached to polymeric micelles for active targeting. 114–117<br />

Park et al. developed a folate-mediated (FOL-mediated) intracellular delivery system that can<br />

transport therapeutic proteins or other bioactive macromolecules into specific cells that over<br />

express FOL receptors. 118,119 The di-block copolymer conjugate, FOL–PEO-b-P(L-Lys)<br />

(Figure 18.8), was physically complexed with fluorescein isothiocyanate conjugated bovine<br />

serum albumin (FITC–BSA) in an aqueous phase <strong>by</strong> ionic interactions. Cellular uptake of<br />

FOL–PEO-b-P(L-Lys)/FITC–BSA complexes was greatly enhanced against a FOL receptor over<br />

expressing cell line (KB cells) compared to a FOL receptor deficient cell line (A549 cells). The<br />

presence of an excess amount of free FOL in the medium inhibited the intracellular delivery of<br />

FOL–PEO-b-P(L-Lys)/FITC–BSA complexes. Therefore, the enhanced cellular uptake of<br />

FITC–BSA <strong>by</strong> KB cells was attributed to FOL receptor-mediated endocytosis of the complexes<br />

having FOL moieties on the surface. The FOL–PEO-b-P(L-Lys) di-block copolymer can potentially<br />

be applied for intracellular delivery of a wide range of other active agents that carry<br />

negative charge.<br />

In a separate research, the novel pH-sensitive, FOL-modified polymeric mixed micelles<br />

composed of FOL–PEO-b-P(L-His) and PEO-b-poly(L-lactic acid) block copolymers were<br />

prepared. The anti-cancer drug, DOX, was physically encapsulated in the micelles. 120 The polymeric<br />

micellar carrier was shown to be more efficient for the delivery of DOX to tumor cells<br />

in vitro, demonstrating the potential for solid tumor treatment through combining targetability<br />

and pH sensitivity.<br />

Attachment of C225, i.e., the antibody against epidermal growth factor (EGF) receptors, to the<br />

PEO terminus of a PEO-b-P(L-Glu)-DOX has been reported (Figure 18.9). 121 The polymeric<br />

immuno-conjugate C225–PEO-b-P(L-Glu)-DOX selectively bound to human vulvar squamous<br />

carcinoma (A431) cells that over express EGF receptors. Receptor-mediated uptake of C225–<br />

PEO-b-P(L-Glu)-DOX occurred rapidly (within 5 min), but non-specific uptake of PEO-b-<br />

P(L-Glu)-DOX required 24 h. Binding of C225–PEO-b-P(L-Glu)-DOX to A431 cells was<br />

blocked <strong>by</strong> pretreatment with C225 antibody. The results indicate that C225–PEO-b-P(L-Glu)-<br />

DOX was more potent than free DOX in inhibiting the growth of A431 cells after a 6-h<br />

exposure period.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


374<br />

H 2 N<br />

H 2 N<br />

N<br />

H<br />

N<br />

O<br />

N<br />

N<br />

FOL<br />

H<br />

N<br />

FOL−PEO-P(L-Lys)-CBZ<br />

N<br />

H<br />

N<br />

O<br />

N<br />

N<br />

HBr/OHAc<br />

H<br />

N H N<br />

O<br />

DCC/NHS<br />

H<br />

N COOH<br />

18.8 PEO-b-PLAA MICELLES FOR GENE DELIVERY<br />

O<br />

COOH<br />

COOH<br />

O<br />

C<br />

FOL-PEG-P(L-Lys)<br />

+<br />

Anti-cancer drugs<br />

ε-CBZ-P(L-Lys)<br />

N<br />

H<br />

FOL−NHS<br />

HOOC−PEO-NH 2<br />

Polymer-based systems have attracted great interest for gene delivery because compared to viral<br />

vectors, they are simple to prepare, rather stable, easy to modify, and relatively safe. For the<br />

purpose of gene delivery, the polymeric carriers should bear positive charges that can interact<br />

with the negative charges of phosphate groups on DNA, resulting in condensation of DNA <strong>by</strong><br />

the carrier. 122 The polymer/DNA complexes are usually prepared in the presence of an excess<br />

amount of cationic polymer, leading to a net positive charge for the complex. The net positive<br />

charge of the complex will increase the chance of its interaction with negatively charged cell<br />

membranes and facilitates cellular uptake of the polymer/DNA complex via endocytosis. P(L-<br />

Lys) and its derivatives have widely been investigated for gene delivery because of its high<br />

charge density. However, the high charge density of P(L-Lys) also results in its cytotoxicity and<br />

its rapid clearance from circulation after intravenous injection. PEO has been attached to P(L-Lys)<br />

to shield the high charge density of the P(L-Lys), reducing its toxicity and elimination from the<br />

body. PEO-P(L-Lys) copolymers are capable of association and micelle formation. To date, two<br />

types of PEO-P(L-Lys) copolymers have been used as gene carriers: AB type block copolymers, i.e.,<br />

PEO-b-P(L-Lys) and comb-shaped PEO grafted P(L-Lys), i.e., PEO-g-P(L-Lys) (Figure 18.10).<br />

The PEO-b-P(L-Lys) was synthesized and investigated as a gene carrier <strong>by</strong> Seymour et al. It<br />

formed carriers with an average diameter of around 100 nm. 123 The cytotoxicity of the PEO-b-<br />

P(L-Lys)/DNA complex was found to be significantly lower than P(L-Lys)/DNA complex. PEO-b-<br />

O n<br />

Nanotechnology for Cancer Therapy<br />

O<br />

FOL−PEO−COOH<br />

O<br />

O<br />

O<br />

N<br />

H CH<br />

C<br />

n<br />

CH2 H2C CH2 H2C NH2 FIGURE 18.8 Synthesis of FOL–PEO-b-P(L-Lys) and formation of polyion complex micelles with charged<br />

anti-cancer drugs or genes.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

OH


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 375<br />

O<br />

S PEO C<br />

O<br />

VS-PEO-NHS<br />

C225-SH<br />

S<br />

SH<br />

+<br />

O<br />

O<br />

O N<br />

PEO<br />

O<br />

O<br />

S PEO C<br />

O<br />

O<br />

S PEO C<br />

O<br />

C225−PEO-P(L-Glu)-DOX<br />

O<br />

H<br />

N<br />

CO CH<br />

CH2 NH<br />

n<br />

CH 2<br />

CH 2<br />

CH 2<br />

NH 2<br />

O<br />

PEO-b-P(L-lysine)<br />

P(L-Glu), PBS<br />

pH 8<br />

H<br />

N<br />

VS-PEO-P(L-Glu)-DOX<br />

H<br />

C C H O<br />

N<br />

CO CH NH n<br />

CH 2<br />

CH 2<br />

CH 2<br />

CH 2<br />

NH<br />

PEO<br />

PEO-g-P(L-lysine)<br />

FIGURE 18.10 Chemical structure of PEO-b-P(L-Lys) 123–127 and PEO-g-P(L-Lys) 129 used for gene delivery.<br />

CH 2<br />

CH 2<br />

n<br />

O C NH DOX<br />

H<br />

C C H O<br />

N<br />

CH2<br />

CH2<br />

O C NH DOX<br />

n<br />

O<br />

O<br />

S PEO C<br />

O<br />

H<br />

N CH<br />

O<br />

C NH<br />

DMF<br />

CH 2<br />

CH 2<br />

C O<br />

OH<br />

Disopropylcabodiimide<br />

FIGURE 18.9 Synthesis of C225–PEO-b-P(L-Glu)-DOX and micelles with chemically loaded anti-cancer<br />

drugs.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

DOX<br />

n


376<br />

P(L-Lys)s were further studied <strong>by</strong> Kataoka et al. for DNA delivery. 124–127 Complexation of DNA<br />

and PEO-b-P(L-Lys) in their study led to the formation of polyion complex micelles with an<br />

average diameter of 48.5 nm. The polyion complex micelles showed higher transfection efficiency<br />

in human hepatoma HepG2 cells at a 4:1 positive to negative charge ratio when compared to<br />

P(L-Lys) of the same molecular weight. Southern blotting assay showed that naked plasmid<br />

DNA was degraded in the blood within 5 min after intravenous injection. On the contrary, when<br />

plasmid DNA was administered as complex with PEO-b-P(L-Lys) micelles, the supercoiled DNA<br />

was detected in the blood for 30 min. Methoxy PEO-b-P(L-Lys) dendrimers have also been studied<br />

as gene carriers <strong>by</strong> Choi et al., forming a spherical polymer/DNA complex at a 2:1 positive to<br />

negative charge ratio. 63 Multiblock copolymers of PEO-b-P(L-Lys)-g-P(L-His) with ester bonds<br />

between PEO and P(L-Lys) have also been investigated as gene carriers. 128<br />

The application of PEO-g-P(L-Lys) as gene carriers was evaluated in vitro and in vivo <strong>by</strong> Kim<br />

et al. PEO-g-P(L-Lys) showed a slightly lower DNA condensing effect when compared to P(L-Lys),<br />

but it had a 5–30-fold increase in transfection efficiency in HepG2 cells (a human liver carcinoma<br />

cell line). 129 PEO-g-P(L-Lys) demonstrated low cytotoxicity, early gene expression, and maintenance<br />

of gene expression level for up to 96 h. PEO-g-P(L-Lys) was then evaluated as a carrier of the<br />

anti-sense glutamic acid decarboxylase (GAD) plasmid to the pancreas for the prevention of type 1<br />

diabetes in mice. 130 The anti-sense mRNA was expressed in the pancreas of mouse for more than 3<br />

days when delivered as PEO-g-P(L-Lys) complex.<br />

Attachment of galactose or lactose to PEO-b-P(L-Lys) nanocarriers has been accomplished to<br />

target the asialoglycoprotein receptor of hepatocytes. Lactose-PEO-b-P(L-Lys) showed ten times<br />

higher transfection efficiency in HepG2 and lower cytotoxicity when compared to P(L-Lys)/DNA<br />

complex. 131<br />

A synthetic peptide based on apolipoprotein B100 with arterial wall-binding effects was<br />

selected and introduced to the end of PEO-g-P(L-Lys) for targeting the polymer/DNA complex<br />

to the aorta endothelial cells. 132 The arterial wall binding (AWBP) conjugate of PEO-g-P(L-Lys)<br />

condensed plasmid DNA and formed a spherical shape complex with a size of 100 nm. In gene<br />

expression studies, the transfection efficiency of the AWBP–PEO-g-P(L-Lys)/DNA complex to<br />

bovine aorta endothelial cells and smooth muscle cells was 150–180 times higher than that of<br />

P(L-Ly) or PEO-g-P(L-Lys) complexes with DNA. Free AWBP decreased the transfection efficiency<br />

of the AWBP–PEO-g-P(L-Lys), suggesting a targeted gene delivery <strong>by</strong> the carrier to the<br />

aorta endothelial cells via receptor-mediated endocytosis.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Aliabadi, H. M. and Lavasanifar, A., Polymeric micelles for drug delivery, Expert Opinion on Drug<br />

Delivery, 3(1), 139–162, <strong>2006</strong>.<br />

2. Gaucher, G., Dufresne, M. H., Sant, V. P., Kang, N., Maysinger, D., and Leroux, J., Block copolymer<br />

micelles: Preparation, characterization and application in drug delivery, Journal of Controlled<br />

Release, 109(1–3), 169–188, 2005.<br />

3. Kwon, G. S., Polymeric micelles for delivery of poorly water-soluble compounds, Critical Reviews<br />

in Therapeutic Drug Carrier Systems, 20(5), 357–403, 2003.<br />

4. Lavasanifar, A., Samuel, J., and Kwon, G. S., Poly(ethylene oxide)-block-poly(L-amino acid)<br />

micelles for drug delivery, Advanced Drug Delivery Reviews, 54(2), 169–190, 2002.<br />

5. Allen, C., Eisenberg, A., and Maysinger, D., Copolymer drug carriers: Conjugates, micelles and<br />

microspheres, STP Pharma Sciences, 9(1), 139–151, 1999.<br />

6. Allen, C., Maysinger, D., and Eisenberg, A., Nano-engineering block copolymer aggregates for drug<br />

delivery, Colloids and Surfaces B: Biointerfaces, 16(1–4), 3–27, 1999.<br />

7. Kataoka, K., Harada, A., and Nagasaki, Y., Block copolymer micelles for drug delivery: Design,<br />

characterization and biological significance, Advanced Drug Delivery Reviews, 47(1), 113–131,<br />

2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 377<br />

8. Le Garrec, D., Ranger, M., and Leroux, J. C., Micelles in anticancer drug delivery, American Journal<br />

of Drug Delivery, 2(1), 15–42, 2004.<br />

9. Jones, M. C. and Leroux, J. C., Polymeric micelles—A new generation of colloidal drug carriers,<br />

European Journal of Pharmaceutics and Biopharmaceutics, 48(2), 101–111, 1999.<br />

10. Kwon, G. S. and Kataoka, K., Block copolymer micelles as long-circulating drug vehicles, Advanced<br />

Drug Delivery Reviews, 16(2–3), 295–309, 1995.<br />

11. Torchilin, V. P., Structure and design of polymeric surfactant-based drug delivery systems, Journal<br />

of Controlled Release, 73(2–3), 137–172, 2001.<br />

12. Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Improved synthesis of adriamycin-conjugated<br />

poly(ethylene oxide)-poly(aspartic acid) block copolymer and formation of unimodal micellar<br />

structure with controlled amount of physically entrapped adriamycin, Journal of Controlled Release,<br />

32(3), 269–277, 1994.<br />

13. Kwon, G. S., Naito, M., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Physical entrapment<br />

of adriamycin in AB block copolymer micelles, Pharmaceutical Research, 12(2), 192–195, 1995.<br />

14. Kwon, G., Naito, M., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Block copolymer<br />

micelles for drug delivery: Loading and release of doxorubicin, Journal of Controlled Release,<br />

48(2–3), 195–201, 1997.<br />

15. Kataoka, K., Matsumoto, T., Yokoyama, M., Okano, T., Sakurai, Y., Fukushima, S., Okamoto, K.,<br />

and Kwon, G. S., Doxorubicin-loaded poly(ethylene glycol)-poly(beta-benzyl-L-aspartate) copolymer<br />

micelles: Their pharmaceutical characteristics and biological significance, Journal of<br />

Controlled Release, 64(1–3), 143–153, 2000.<br />

16. Yokoyama, M., Fukushima, S., Uehara, R., Okamoto, K., Kataoka, K., Sakurai, Y., and Okano, T.,<br />

Characterization of physical entrapment and chemical conjugation of adriamycin in polymeric<br />

micelles and their design for in vivo delivery to a solid tumor, Journal of Controlled Release,<br />

50(1–3), 79–92, 1998.<br />

17. Yokoyama, M., Okano, T., Sakurai, Y., Fukushima, S., Okamoto, K., and Kataoka, K., Selective<br />

delivery of adriamycin to a solid tumor using a polymeric micelle carrier system, Journal of Drug<br />

Targeting, 7(3), 171–186, 1999.<br />

18. Kwon, G. S., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Biodistribution of micelleforming<br />

polymer-drug conjugates, Pharmaceutical Research, 10(7), 970–974, 1993.<br />

19. Yokoyama, M., Miyauchi, M., Yamada, N., Okanao, T., Sakurai, Y., Kataoka, K., and Inoue, S.,<br />

Polymer micelles as novel drug carrier: Adriamycin-conjugated poly(ethylene glycol)–poly(aspartic<br />

acid) block copolymer, Journal of Controlled Release, 11(1–3), 269–278, 1990.<br />

20. Kwon, G., Suwa, S., Yokayama, M., Okano, T., Sakurai, Y., and Kataoka, K., Enhanced tumor<br />

accumulation and prolonged circulation times for micelle-forming poly(ethylene oxide-aspartate)<br />

block copolymer–adriamycin conjugates, Journal of Controlled Release, 29, 17–23, 1994.<br />

21. Nakanishi, T., Fukushima, S., Okamoto, K., Suzuki, M., Matsumura, Y., Yokoyama, M., Okano, T.,<br />

Sakurai, Y., and Kataoka, K., Development of the polymer micelle carrier system for doxorubicin,<br />

Journal of Controlled Release, 74(1–3), 295–302, 2001.<br />

22. Yokoyama, M., Okano, T., Sakurai, Y., Ekimoto, H., Shibazaki, C., and Kataoka, K., Toxicity and<br />

antitumor activity against solid tumors of micelle-forming polymeric anticancer drug and its extremely<br />

long circulation in blood, Cancer Research, 51(12), 3229–3236, 1991.<br />

23. Matsumura, Y., Hamaguchi, T., Ura, T., Muro, K., Yamada, Y., Shimada, Y., Shirao, K., Okusaka,<br />

T., Ueno, H., Ikeda, M., and Watanabe, N., Phase I clinical trial and pharmacokinetic evaluation of<br />

NK911, a micelle-encapsulated doxorubicin, British Journal of Cancer, 91(10), 1775–1781, 2004.<br />

24. Hamaguchi, T., Matsumura, Y., Suzuki, M., Shimizu, K., Goda, R., Nakamura, I., Nakatomi, I.,<br />

Yokoyama, M., Kataoka, K., and Kakizoe, T., NK105, a paclitaxel-incorporating micellar nanoparticle<br />

formulation, can extend in vivo antitumour activity and reduce the neurotoxicity of paclitaxel,<br />

British Journal of Cancer, 92, 1240–1246, 2005.<br />

25. Bogdanov, A. A., Martin, C., Bogdanova, A. V., Brady, T. J., and Weissleder, R., An adduct of cisdiamminedichloroplatinum(II)<br />

and poly(ethylene glycol)poly(L-lysine)-succinate: Synthesis and<br />

cytotoxic properties, Bioconjugate Chemistry, 7(1), 144–149, 1996.<br />

26. Nishiyama, N. and Kataoka, K., Preparation and characterization of size-controlled polymeric<br />

micelle containing cis-dichlorodiammineplatinum(II) in the core, Journal of Controlled Release,<br />

74(1–3), 83–94, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


378<br />

Nanotechnology for Cancer Therapy<br />

27. Nishiyama, N., Kato, Y., Sugiyama, Y., and Kataoka, K., Cisplatin-loaded polymer–metal complex<br />

micelle with time-modulated decaying property as a novel drug delivery system, Pharmaceutical<br />

Research, 18(7), 1035–1041, 2001.<br />

28. Nishiyama, N., Koizumi, F., Okazaki, S., Matsumura, Y., Nishio, K., and Kataoka, K., Differential<br />

gene expression profile between PC-14 cells treated with free cisplatin and cisplatin-incorporated<br />

polymeric micelles, Bioconjugate Chemistry, 14(2), 449–457, 2003.<br />

29. Nishiyama, N., Okazaki, S., Cabral, H., Miyamoto, M., Kato, Y., Sugiyama, Y., Nishio, K.,<br />

Matsumura, Y., and Kataoka, K., Novel cisplatin-incorporated polymeric micelles can eradicate<br />

solid tumors in mice, Cancer Research, 63(24), 8977–8983, 2003.<br />

30. Nishiyama, N., Yokoyama, M., Aoyaga, T., Okano, T., Sakurai, Y., and Kataoka, K., Preparation and<br />

characterization of self-assembled polymer–metal complex micelle from cis-dichlorodiammineplatinium<br />

(II) and poly(ethylene glycole)–poly(aspartic acid) block copolymer in an aqueous medium,<br />

Langmuir, 15, 377–383, 1999.<br />

31. Mizumura, Y., Matsumura, Y., Hamaguchi, T., Nishiyama, N., Kataoka, K., Kawaguchi, T.,<br />

Hrushesky, W. J., Moriyasuand, F., and Kakizoe, T., Cisplatin-incorporated polymeric micelles<br />

eliminate nephrotoxicity, while maintaining antitumor activity, Japanese Journal of Cancer<br />

Research, 92(3), 328–336, 2001.<br />

32. Uchino, H., Matsumura, Y., Negishi, T., Koizumi, F., Hayashi, T., Honda, T., Nishiyama, N.,<br />

Kataoka, K., Naito, S., and Kakizoe, T., Cisplatin-incorporating polymeric micelles (NC-6004)<br />

can reduce nephrotoxicity and neurotoxicity of cisplatin in rats, British Journal of Cancer, 93(6),<br />

678–687, 2005.<br />

33. Adams, M. L., Andes, D. R., and Kwon, G. S., Amphotericin B encapsulated in micelles based on<br />

poly(ethylene oxide)-block-poly(L-amino acid) derivatives exerts reduced in vitro hemolysis but<br />

maintains potent in vivo antifungal activity, Biomacromolecules, 4(3), 750–757, 2003.<br />

34. Adams, M. L. and Kwon, G. S., Relative aggregation state and hemolytic activity of amphotericin B<br />

encapsulated <strong>by</strong> poly(ethylene oxide)-block-poly(N-hexyl-L-aspartamide)-acyl conjugate micelles:<br />

Effects of acyl chain length, Journal of Controlled Release, 87(1–3), 23–32, 2003.<br />

35. Lavasanifar, A., Samuel, J., and Kwon, G. S., Micelles of poly(ethylene oxide)-block-poly(N-alkyl<br />

stearate L-aspartamide): Synthetic analogues of lipoproteins for drug delivery, Journal of Biomedical<br />

Materials Research, 52(4), 831–835, 2000.<br />

36. Lavasanifar, A., Samuel, J., and Kwon, G. S., Micelles self-assembled from poly(ethylene oxide)block-poly(N-hexyl<br />

stearate L-aspartamide) <strong>by</strong> a solvent evaporation method: Effect on the solubilization<br />

and haemolytic activity of amphotericin B, Journal of Controlled Release, 77(1–2), 155–160,<br />

2001.<br />

37. Lavasanifar, A., Samuel, J., and Kwon, G. S., The effect of alkyl core structure on micellar properties<br />

of poly(ethylene oxide)-block-poly(L-aspartamide) derivatives, Colloids and Surfaces B: Biointerfaces,<br />

22(2), 115–126, 2001.<br />

38. Lavasanifar, A., Samuel, J., and Kwon, G. S., The effect of fatty acid substitution on the in vitro<br />

release of amphotericin B from micelles composed of poly(ethylene oxide)-block-poly(N-hexyl<br />

stearate-L-aspartamide), Journal of Controlled Release, 79(1–3), 165–172, 2002.<br />

39. Lavasanifar, A., Samuel, J., Sattari, S., and Kwon, G. S., Block copolymer micelles for the encapsulation<br />

and delivery of amphotericin B, Pharmaceutical Research, 19(4), 418–422, 2002.<br />

40. Yu, B. G., Okano, T., Kataoka, K., and Kwon, G. S., Polymeric micelles for drug delivery: Solubilization<br />

and haemolytic activity of amphotericin B, Journal of Controlled Release, 53(1–3),<br />

131–136, 1998.<br />

41. Yu, B. G., Okano, T., Kataoka, K., and Kwon, G. S., In vitro dissociation of antifungal efficacy and<br />

toxicity for amphotericin B-loaded poly(ethylene oxide)-block-poly(beta-benzyl-L-aspartate)<br />

micelles, Journal of Controlled Release, 56(1–3), 285–291, 1998.<br />

42. Smeenk, J. M., Lowik, D. W. P. M., and van Hest, J. C. M., Peptide-containing block copolymers:<br />

Synthesis and potential applications of bio-mimetic materials, Current Organic Chemistry, 9(12),<br />

1115–1125, 2005.<br />

43. Fuller, W. D., Verlander, M. S., and Goodman, M., A procedure for the facile synthesis of aminoacid<br />

N-carboxyanhydrides, Biopolymers, 15(9), 1869–1871, 1976.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 379<br />

44. Li, L. Y., Sun, P. C., Yao, Y., Chen, T. H., Li, B. H., Jin, Q. H., and Ding, D. T., Synthesis of<br />

amphiphilic diblock copolymer poly(L-alanine)-b-poly(hydroxyethyl glutamine) and its selfassembly<br />

in water, Chemical Journal of Chinese Universities—Chinese, 26(8), 1548–1551, 2005.<br />

45. Harwood, H. J., Comments concerning the mechanism of strong base initiated NCA polymerization.<br />

Abstract in Abstracts of papers of the American Chemical Society, 187(APR),72-Poly, 1984.<br />

46. Kricheldorf, H. R. and Mulhaupt, R., Mechanism of the NCA polymerization. 7. Primary and<br />

secondary amine-initiated polymerization of beta-amino acid NCAs, Makromolekulare Chemie-<br />

Macromolecular Chemistry and Physics, 180(6), 1419–1433, 1979.<br />

47. Kricheldorf, H. R., Von Lossow, C., and Schwarz, G., Primary amine and solvent-induced polymerizations<br />

of L- or D,L-phenylalanine N-carboxyanhydride, Macromolecular Chemistry and Physics,<br />

206(2), 282–290, 2005.<br />

48. Sekiguchi, H., Mechanism of N-carboxy-alpha-amino acid anhydride (NCA) polymerization, Pure<br />

and Applied Chemistry, 53(9), 1689–1714, 1981.<br />

49. Harada, A. and Kataoka, K., Formation of polyion complex micelles in an aqueous milieu from a pair<br />

of oppositely-charged block-copolymers with poly(ethylene glycol) segments, Macromolecules,<br />

28(15), 5294–5299, 1995.<br />

50. Deming, T. J., Methodologies for preparation of synthetic block copolypeptides: Materials with<br />

future promise in drug delivery, Advanced Drug Delivery Reviews, 54(8), 1145–1155, 2002.<br />

51. Deming, T. J., Transition metal-amine initiators for preparation of well-defined poly(gamma-benzyl<br />

L-glutamate), Journal of the American Chemical Society, 119(11), 2759–2760, 1997.<br />

52. Freireic, S., Gertner, D., and Zilkha, A., Polymerization of N-carboxy anhydrides <strong>by</strong> organotin<br />

catalysts, European Polymer Journal, 10(5), 439–443, 1974.<br />

53. Tsuruta, T., Matsuura, K., and Inoue, S., Copolymerization of propylene oxide with N-carboxy-DLalanine<br />

anhydride <strong>by</strong> organometallic systems, Die Makromolekulare Chemie, 83, 289–291, 1965.<br />

54. Yamashit, S. and Tani, H., Polymerization of gamma-benzyl L-glutamate N-carboxyanhydride with<br />

metal acetate-tri-normal-butylphosphine catalyst system, Macromolecules, 7(4), 406–409, 1974.<br />

55. Yamashit, S., Waki, K., Yamawaki, N., and Tani, H., Stereoselective polymerization of alphaamino-acid<br />

N-carboxyanhydrides with nickel DL-2-methylbutyrate-tri-normal-butylphosphine catalyst<br />

system, Macromolecules, 7(4), 410–415, 1974.<br />

56. Brzezinska, K. R. and Deming, T. J., Synthesis of AB diblock copolymers <strong>by</strong> atom-transfer radical<br />

polymerization (ATRP) and living polymerization of alpha-amino acid-N-carboxyanhydrides,<br />

Macromolecular Bioscience, 4(6), 566–569, 2004.<br />

57. Dvorak, M. and Rypacek, F., Preparation and polymerization of N-carboxyanhydrides of alphaamino-acids,<br />

Chemicke Listy, 89(7), 423–436, 1995.<br />

58. Vayaboury, W., Giani, O., Cottet, H., Deratani, A., and Schue, F., Living polymerization of alphaamino<br />

acid N-carboxyanhydrides (NCA) upon decreasing the reaction temperature, Macromolecular<br />

Rapid Communications, 25(13), 1221–1224, 2004.<br />

59. Deming, T. J., Living polymerization of alpha-amino acid-N-carboxyanhydrides, Journal of Polymer<br />

Science Part A—Polymer Chemistry, 38(17), 3011–3018, 2000.<br />

60. Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., Poly(L-histidine)–PEG block copolymer micelles and<br />

pH-induced destabilization, Journal of Controlled Release, 90(3), 363–374, 2003.<br />

61. Won, C. Y., Chu, C. C., and Lee, J. D., Synthesis and characterization of biodegradable poly(Laspartic<br />

acid-co-PEG), Journal of Polymer Science Part A—Polymer Chemistry, 36(16), 2949–2959,<br />

1998.<br />

62. Van Domeselaar, G. H., Kwon, G. S., Andrew, L. C., and Wishart, D. S., Application of solid phase<br />

peptide synthesis to engineering PEO-peptide block copolymers for drug delivery, Colloids and<br />

Surfaces B: Biointerfaces, 30(4), 323–334, 2003.<br />

63. Choi, J. S., Lee, E. J., Choi, Y. H., Jeong, Y. J., and Park, J. S., Poly(ethylene glycol)-block-poly(Llysine)<br />

dendrimer: Novel linear polymer/dendrimer block copolymer forming a spherical watersoluble<br />

polyionic complex with DNA, Bioconjugate Chemistry, 10(1), 62–65, 1999.<br />

64. La, S. B., Okano, T., and Kataoka, K., Preparation and characterization of the micelle-forming<br />

polymeric drug indomethacin-incorporated poly(ethylene oxide)–poly(beta-benzyl L-aspartate)<br />

block copolymer micelles, Journal of Pharmaceutical Sciences, 85(1), 85–90, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


380<br />

Nanotechnology for Cancer Therapy<br />

65. Yokoyama, M., Satoh, A., Sakurai, Y., Okano, T., Matsumura, Y., Kakizoe, T., and Kataoka, K.,<br />

Incorporation of water-insoluble anticancer drug into polymeric micelles and control of their particle<br />

size, Journal of Controlled Release, 55(2–3), 219–229, 1998.<br />

66. Opanasopit, P., Yokoyama, M., Watanabe, M., Kawano, K., Maitani, Y., and Okano, T., Influence of<br />

serum and albumins from different species on stability of camptothecin-loaded micelles, Journal of<br />

Controlled Release, 104(2), 313–321, 2005.<br />

67. Working, P., Newman, M., Johnson, J., and Cornacoff, J., Safety of poly(ethylene glycol) and<br />

poly(ethylene glycol) derivatives, In Poly(ethylene glycol) Chemistry and Biological Applications,<br />

Harris, J. and Zalipsky, S., Eds., American Chemical Society, Washington, DC, pp. 44–57, 1997.<br />

68. Zalipsky, S. and Harris, J., Introduction to chemistry and biological applications of poly(ethylene<br />

glycol), In Poly(ethylene glycol) Chemistry and Biological Applications, Harris, J. and Zalipsky, S.,<br />

Eds., American Chemical Society, Washington, DC, pp. 1–13, 1997.<br />

69. Stolnik, S., Illum, L., and Davis, S. S., Long circulating microparticulate drug carriers, Advanced<br />

Drug Delivery Reviews, 16, 195–214, 1995.<br />

70. Zhang, F., Kang, E., Neoh, K., and Huang, W., Modification of gold surface <strong>by</strong> grafting of poly(ethylene<br />

glycol) for reduction in protein adsorption and platelet adhesion, Journal of Biomaterial<br />

Science Polymer Edition, 12, 515–531, 2001.<br />

71. Trubetskoy, V. and Torchilin, V., Use of polyoxyethylene-lipid conjugates as long-circulating<br />

carriers for delivery of therapeutic and diagnostic agents, Advanced Drug Delivery Reviews, 16,<br />

311–320, 1995.<br />

72. Dunn, S., Brindley, A., Davis, S., Davies, M., and Illum, L., Polystyrene-poly(ethylene glycol) (PS-<br />

PEG) particles as model systems for site specific drug delivery. 2. The effect of PEG surface density<br />

on the in vitro cell interaction and in vivo biodistribution, Pharmaceutical Research, 11, 1016–1022,<br />

1994.<br />

73. Mosqueira, V. C., Legrand, P., Morgat, J. L., Vert, M., Mysiakine, E., Gref, R., Devissaguet, J. P.,<br />

and Barratt, G., Biodistribution of long-circulating PEG-grafted nanocapsules in mice: Effects of<br />

PEG chain length and density, Pharmaceutical Research, 18(10), 1411–1419, 2001.<br />

74. Gref, R., Luck, M., Quellec, P., Marchand, M., Dellacherie, E., Harnisch, S., Blunk, T., and Muller,<br />

R. H., ‘Stealth’ corona-core nanoparticles surface modified <strong>by</strong> polyethylene glycol (PEG): Influences<br />

of the corona (PEG chain length and surface density) and of the core composition on<br />

phagocytic uptake and plasma protein adsorption, Colloids and Surfaces B: Biointerfaces,<br />

18(3–4), 301–313, 2000.<br />

75. Mahmud, A. and Lavasanifar, A., The effect of block copolymer structure on the internalization of<br />

polymeric micelles <strong>by</strong> human breast cancer cells, Colloids and Surfaces B: Biointerfaces, 45(2),<br />

82–89, 2005.<br />

76. Nishiyama, N., Bae, Y., Miyata, K., Fukushima, S., and Kataoka, K., Smart polymeric micelles for<br />

gene and drug delivery, Drug Discovery Today: Technologies, 2(1), 21–26, 2005.<br />

77. McCormick-Thomson, L., Sgouras, D., and Duncan, R., Poly(amino acid) copolymers as potential<br />

soluble drug delivery system. 2. Body distribution and preliminary biocompatibility testing in vitro<br />

and in vivo, Journal of Bioactive and Compatible Polymers, 4, 252–268, 1989.<br />

78. Chiu, H., Kopeckova, P., and Deshmane, A. S. K. J., Lysosomal degradability of poly(a-amino<br />

acids), Journal of Biomedical Materials Research, 34, 381–392, 1997.<br />

79. Kwon, G. S., Diblock copolymer nanoparticles for drug delivery, Critical Reviews in Therapeutic<br />

Drug Carrier Systems, 15(5), 481–512, 1998.<br />

80. Litzinger, D. C., Buiting, A. M., van Rooijen, N., and Huang, L., Effect of liposome size on the<br />

circulation time and intraorgan distribution of amphipathic poly(ethylene glycol)-containing liposomes,<br />

Biochimica et Biophysica Acta, 1190, 99–107, 1994.<br />

81. Papahadjopoulos, D. and Gabizon, A., Liposomes designed to avoid the reticuloendothelial system,<br />

Progress in Clinical and Biological Research, 343, 85–93, 1990.<br />

82. Charrois, G. J. R. and Allen, T. M., Rate of biodistribution of STEALTH liposomes to tumor and<br />

skin: Influence of liposome diameter and implications for toxicity and therapeutic activity, Biochimica<br />

et Biophysica Acta, 1609(1), 102–108, 2003.<br />

83. Jain, R. K., Delivery of molecular and cellular medicine to solid tumors, Advanced Drug Delivery<br />

Reviews, 46(1–3), 149–168, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 381<br />

84. Dass, C., Tumour angiogenesis, vascular biology and enhanced drug delivery, Journal of Drug<br />

Targeting, 12(5), 245–255, 2004.<br />

85. Duncan, R., The dawning era of polymer therapeutics, Nature Reviews, 2, 347–360, 2003.<br />

86. Maeda, H., Wu, J., Sawa, T., Matsumura, Y., and Hori, K., Tumor vascular permeability and the EPR<br />

effect in macromolecular therapeutics: A review, Journal of Controlled Release, 65(1–2), 271–284,<br />

2000.<br />

87. Li, Y. and Kwon, G. S., Methotrexate esters of poly(ethylene oxide)-block-poly(2-hydroxyethyl-Laspartamide).<br />

Part I: Effects of the level of methotrexate conjugation on the stability of micelles and<br />

on drug release, Pharmaceutical Research, 17(5), 607–611, 2000.<br />

88. Kwon, G., Naito, M., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Micelles based on AB<br />

block copolymers of poly(ethylene oxide) and poly(beta-benzyl L-aspartate), Langmuir, 9, 945–949,<br />

1993.<br />

89. Kakizawa, Y., Harada, A., and Kataoka, K., Glutathione-sensitive stabilization of block copolymer<br />

micelles composed of antisense DNA and thiolated poly(ethylene glycol)-block-poly(L-lysine): A<br />

potential carrier for systemic delivery of antisense DNA, Biomacromolecules, 2(2), 491–497, 2001.<br />

90. Li, Y. and Kwon, G. S., Micelle-like structures of poly(ethylene oxide)-block-poly(2-hydroxyethyl<br />

aspartamide)-methotrexate conjugates, Colloids and Surfaces B: Biointerfaces, 16, 217–226, 1999.<br />

91. Yokoyama, M., Kwon, G. S., Okano, T., Sakurai, Y., Naito, M., and Kataoka, K., Influencing factors<br />

on in vitro micelle stability of adriamycin–block copolymer conjugates, Journal of Controlled<br />

Release, 28(1–3), 59–65, 1994.<br />

92. Jaturanpinyo, M., Harada, A., Yuan, X., and Kataoka, K., Preparation of bionanoreactor based on<br />

core–shell structured polyion complex micelles entrapping trypsin in the core cross-linked with<br />

glutaraldehyde, Bioconjugate Chemistry, 15(2), 344–348, 2004.<br />

93. Miyata, K., Kakizawa, Y., Nishiyama, N., Harada, A., Yamasaki, Y., Koyama, H., and Kataoka, K.,<br />

Block catiomer polyplexes with regulated densities of charge and disulfide cross-linking directed to<br />

enhance gene expression, Journal of the American Chemical Society, 126(8), 2355–2361, 2004.<br />

94. Kataoka, K., Matsumoto, T., Yokoyama, M., Okano, T., Sakurai, Y., Fukushima, S., Okamoto, K.,<br />

and Kwon, G. S., Doxorubicin-loaded poly(ethylene glycol)–poly(beta-benzyl-L-aspartate) copolymer<br />

micelles: Their pharmaceutical characteristics and biological significance, Journal of<br />

Controlled Release, 64(1–3), 143–153, 2000.<br />

95. Kwon, G., Naito, M., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Block copolymer<br />

micelles for drug delivery: Loading and release of doxorubicin, Journal of Controlled Release, 48,<br />

195–201, 1997.<br />

96. Ringsdorf, H., Dorn, K., and Hoerpel, G., Polymeric antitumor agents on molecular and cellular<br />

level, In Bioactive Polymer Systems: An Overview, Gebelein, C. G. and Carraher, C. E., Eds., Plenum<br />

Press, New York, pp. 531–585, 1985.<br />

97. Kwon, G., Suwa, S., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Enhanced tumor<br />

accumulation and prolonged circulation times of micelle-forming poly(ethylene oxide-aspartate)<br />

block copolymer–adriamycin conjugates, Journal of Controlled Release, 29, 17–23, 1994.<br />

98. Yokoyama, M., Inoue, S., Kataoka, K., Yui, N., and Sakurai, Y., Preparation of adriamycin-conjugated<br />

poly(ethylene glycol)–poly(aspartic acid) block copolymer: A new type of polymeric<br />

anticancer agent, Makromolekulare Chemie, 8, 431–435, 1987.<br />

99. Yokoyama, M., Kwon, G. S., Okano, T., Sakurai, Y., Seto, T., and Kataoka, K., Preparation of<br />

micelle forming polymer–drug conjugates, Bioconjugate Chemistry, 3, 295–301, 1992.<br />

100. Yokoyama, M., Miyauchi, M., Yamada, N., Okano, T., Sakurai, Y., and Kataoka, K., Characterization<br />

and anticancer activity of the micelle-forming polymeric anticancer drug adriamycinconjugated<br />

poly(ethylene glycol)–poly(aspartic acid) block copolymer, Cancer Research, 50,<br />

1693–1700, 1990.<br />

101. Kim, T. Y., Kim, D. W., Chung, J. Y., Shin, S. G., Kim, S. C., Heo, D. S., Kim, N. K., and Bang,<br />

Y. J., Phase I and pharmacokinetic study of Genexol-PM, a cremophor-free, polymeric micelleformulated<br />

paclitaxel, in patients with advanced malignancies, Clinical Cancer Research, 10(11),<br />

3708–3716, 2004.<br />

102. Kim, S. C., Kim, D. W., Shim, Y. H., Bang, J. S., Oh, H. S., Wan Kim, S., and Seo, M. H., In vivo<br />

evaluation of polymeric micellar paclitaxel formulation: Toxicity and efficacy, Journal of Controlled<br />

Release, 72(1–3), 191–202, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


382<br />

Nanotechnology for Cancer Therapy<br />

103. Le Garrec, D., Gori, S., Luo, L., Lessard, D., Smith, D. C., Yessine, M. A., Ranger, M., and Leroux,<br />

J. C., Poly(N-vinylpyrrolidone)-block-poly(D,L-lactide) as a new polymeric solubilizer for hydrophobic<br />

anticancer drugs: In vitro and in vivo evaluation, Journal of Controlled Release, 99(1),<br />

83–101, 2004.<br />

104. Leung, S. Y., Jackson, J., Miyake, H., Burt, H., and Gleave, M. E., Polymeric micellar paclitaxel<br />

phosphorylates Bcl-2 and induces apoptotic regression of androgen-independent LNCaP prostate<br />

tumors, Prostate, 44(2), 156–163, 2000.<br />

105. Zhang, X., Burt, H. M., Mangold, G., Dexter, D., Von Hoff, D., Mayer, L., and Hunter, W. L., Antitumor<br />

efficacy and biodistribution of intravenous polymeric micellar paclitaxel, Anti-Cancer Drugs,<br />

8(7), 696–701, 1997.<br />

106. Zhang, X., Burt, H. M., Von Hoff, D., Dexter, D., Mangold, G., Degen, D., Oktaba, A. M., and<br />

Hunter, W. L., An investigation of the antitumour activity and biodistribution of polymeric micellar<br />

paclitaxel, Cancer Chemotherapy and Pharmacology, 40(1), 81–86, 1997.<br />

107. Zhang, X., Jackson, J. K., and Burt, H. M., Development of amphiphilic diblock copolymers as<br />

micellar carriers of taxol, International Journal of Pharmaceutics, 132(1–2), 195–206, 1996.<br />

108. Cabral, H., Nishiyama, N., Okazaki, S., Koyama, H., and Kataoka, K., Preparation and biological<br />

properties of dichloro(1,2-diaminocyclohexane)platinum(II) (DACHPt)-loaded polymeric micelles,<br />

Journal of Controlled Release, 101(1–3), 223–232, 2005.<br />

109. Sakai, T., Kawai, H., Kamishohara, M., Odagawa, A., Suzuki, A., Uchida, T., Kawasaki, T., Tsuruo,<br />

T., and Otake, N., Structure–antitumor activity relationship of semi-synthetic spicamycin derivatives,<br />

Journal of Antibiotics (Tokyo), 48(12), 1467–1480, 1995.<br />

110. Matsumura, Y., Yokoyama, M., Kataoka, K., Okano, T., Sakurai, Y., Kawaguchi, T., and Kakizoe,<br />

T., Reduction of the side effects of an antitumor agent, KRN5500, <strong>by</strong> incorporation of the drug into<br />

polymeric micelles, Japanese Journal of Cancer Research, 90(1), 122–128, 1999.<br />

111. Barreiro-Iglesias, R., Bromberg, L., Temchenko, M., Hatton, T. A., Concheiro, A., and Alvarez-<br />

Lorenzo, C., Solubilization and stabilization of camptothecin in micellar solutions of pluronic-gpoly(acrylic<br />

acid) copolymers, Journal of Controlled Release, 97(3), 537–549, 2004.<br />

112. Mi, Z. and Burke, T. G., Differential interactions of camptothecin lactone and carboxylate forms<br />

with human blood components, Biochemistry, 33(34), 10325–10336, 1994.<br />

113. Yokoyama, M., Opanasopit, P., Okano, T., Kawano, K., and Maitani, Y., Polymer design and<br />

incorporation methods for polymeric micelle carrier system containing water-insoluble anticancer<br />

agent camptothecin, Journal of Drug Targeting, 12(6), 373–384, 2004.<br />

114. Leamon, C. P., Weigl, D., and Hendren, R. W., Folate copolymer-mediated transfection of cultured<br />

cells, Bioconjugate Chemistry, 10(6), 947–957, 1999.<br />

115. Nagasaki, Y., Yasugi, K., Yamamoto, Y., Harada, A., and Kataoka, K., Sugar-installed polymeric<br />

micelle for a vehicle of an active targeting drug delivery system, Abstract in Abstracts of Papers of<br />

the American Chemical Society, publication no. 221, U434–U434, 2001.<br />

116. Nasongkla, N., Shuai, X., Ai, H., Weinberg, B. D., Pink, J., Boothman, D. A., and Gao, J. M., cRGDfunctionalized<br />

polymer micelles for targeted doxorubicin delivery, Angewandte Chemie-International<br />

Edition, 43(46), 6323–6327, 2004.<br />

117. Vinogradov, S., Batrakova, E., Li, S., and Kabanov, A., Polyion complex micelles with proteinmodified<br />

corona for receptor-mediated delivery of oligonucleotides into cells, Bioconjugate Chemistry,<br />

10(5), 851–860, 1999.<br />

118. Kim, S. H., Jeong, J. H., Joe, C. O., and Park, T. G., Folate receptor mediated intracellular protein<br />

delivery using PLL–PEG–FOL conjugate, Journal of Controlled Release, 103(3), 625–634, 2005.<br />

119. Kim, S. H., Jeong, J. H., Chun, K. W., and Park, T. G., Target-specific cellular uptake of PLGA<br />

nanoparticles coated with poly(L-lysine)-poly(ethylene glycol)–folate conjugate, Langmuir, 21(19),<br />

8852–8857, 2005.<br />

120. Lee, E. S., Na, K., and Bae, Y. H., Polymeric micelle for tumor pH and folate-mediated targeting,<br />

Journal of Controlled Release, 91(1–2), 103–113, 2003.<br />

121. Vega, J., Ke, S., Fan, Z., Wallace, S., Charsangavej, C., and Li, C., Targeting doxorubicin to<br />

epidermal growth factor receptors <strong>by</strong> site-specific conjugation of C225 to poly(L-glutamic acid)<br />

through a polyethylene glycol spacer, Pharmaceutical Research, 20(5), 826–832, 2003.<br />

122. Kakizawa, Y. and Kataoka, K., Block copolymer micelles for delivery of gene and related<br />

compounds, Advanced Drug Delivery Reviews, 54, 203–222, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEO-Modified Poly(L-Amino Acid) Micelles for Drug Delivery 383<br />

123. Wolfert, M. A., Schacht, E. H., Toncheva, V., Ulbrich, K., Nazarova, O., and Seymour, L. W.,<br />

Characterization of vectors for gene therapy formed <strong>by</strong> self-assembly of DNA with synthetic block<br />

co-polymers, Human Gene Theraphy, 7(17), 2123–2133, 1996.<br />

124. Katayose, S. and Kataoka, K., Remarkable increase in nuclease resistance of plasmid DNA through<br />

supramolecular assembly with poly(ethylene glycol)–poly(L-lysine) block copolymer, Journal of<br />

Pharmaceutical Sciences, 87(2), 160–163, 1998.<br />

125. Katayose, S. and Kataoka, K., Water-soluble polyion complex associates of DNA and poly(ethylene<br />

glycol)–poly(L-lysine) block copolymer, Bioconjugate Chemistry, 8(5), 702–707, 1997.<br />

126. Harada-Shiba, M., Yamauchi, K., Harada, A., Takamisawa, I., Shimokado, K., and Kataoka, K.,<br />

Polyion complex micelles as vectors in gene therapy—pharmacokinetics and in vivo gene transfer,<br />

Gene Theraphy, 9(6), 407–414, 2002.<br />

127. Harada, A., Togawa, H., and Kataoka, K., Physicochemical properties and nuclease resistance of<br />

antisense-oligodeoxynucleotides entrapped in the core of polyion complex micelles composed of<br />

poly(ethylene glycol)–poly(L-lysine) block copolymers, European Journal of Pharmaceutical<br />

Sciences, 13(1), 35–42, 2001.<br />

128. Bikram, M., Ahn, C. H., Chae, S. Y., Lee, M. Y., Yockman, J. W., and Kim, S. W., Biodegradable<br />

poly(ethylene glycol)-co-poly(L-lysine)-g-histidine multiblock copolymers for nonviral gene<br />

delivery, Macromolecules, 37(5), 1903–1916, 2004.<br />

129. Choi, Y. H., Liu, F., Kim, J. S., Choi, Y. K., Park, J. S., and Kim, S. W., Polyethylene glycol-grafted<br />

poly-L-lysine as polymeric gene carrier, Journal of Controlled Release, 54(1), 39–48, 1998.<br />

130. Lee, M., Han, S. O., Ko, K. S., Koh, J. J., Park, J. S., Yoon, J. W., and Kim, S. W., Repression of<br />

GAD autoantigen expression in pancreas beta-cells <strong>by</strong> delivery of antisense plasmid/PEG-g-PLL<br />

complex, Molecular Therapy, 4(4), 339–346, 2001.<br />

131. Choi, Y. H., Liu, F., Park, J. S., and Kim, S. W., Lactose-poly(ethylene glycol)-grafted poly-L-lysine<br />

as hepatoma cell-targeted gene carrier, Bioconjugate Chemistry, 9(6), 708–718, 1998.<br />

132. Nah, J. W., Yu, L., Han, S. O., Ahn, C. H., and Kim, S. W., Artery wall binding peptide–poly(ethylene<br />

glycol)-grafted-poly(L-lysine)-based gene delivery to artery wall cells, Journal of<br />

Controlled Release, 78(1-3), 273–284, 2002.<br />

133. Jeong, Y. I., Nah, J. W., Lee, H. C., Kim, S. H., and Cho, C. S., Adriamycin release from flower-type<br />

polymeric micelle based on star-block copolymer composed of poly(gamma-benzyl L-glutamate) as<br />

the hydrophobic part and poly(ethylene oxide) as the hydrophilic part, International Journal of<br />

Pharmaceutics, 188(1), 49–58, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


19<br />

CONTENTS<br />

Hydrotropic Polymer Micelles<br />

for Cancer Therapeutics<br />

Sang Cheon Lee, Kang Moo Huh, Tooru Ooya, and<br />

Kinam Park<br />

19.1 Introduction ....................................................................................................................... 385<br />

19.2 General Review of Polymer Micelle Drug Carriers ......................................................... 387<br />

19.2.1 Preparative Methods of Polymer Micelles Based on Block Copolymers.......... 387<br />

19.2.2 Characterization and Properties of the Polymeric Micelles............................... 388<br />

19.2.3 Key Parameters for Morphology and Stabilization<br />

of the Polymeric Micelles ................................................................................... 389<br />

19.2.4 Drug Loading, Solubilization, and Drug Release............................................... 390<br />

19.3 Polymer Micelles for Cancer Chemotherapy ................................................................... 391<br />

19.3.1 Polymer Micelles as Carriers of Anti-Cancer Drugs ......................................... 391<br />

19.3.2 Polymer Micelles for Solubilization of Poorly Soluble Anti-Cancer Drugs ..... 391<br />

19.3.3 Targeting Systems Using Polymer Micelles ...................................................... 392<br />

19.3.4 Other Applications of Polymer Micelles for Cancer Therapy ........................... 393<br />

19.4 Hydrotropic Polymer Micelles .......................................................................................... 393<br />

19.4.1 Hydrotropy and Hydrotropic Agents .................................................................. 393<br />

19.4.2 Hydrotropic Agents in Pharmaceutics ................................................................ 394<br />

19.4.3 Mechanistic Studies and Structure–Property Relationship of Hydrotropic<br />

Solubilization....................................................................................................... 395<br />

19.4.4 Design Strategy of Hydrotropic Polymer Micelle Systems ............................... 395<br />

19.4.5 Hydrotropic Polymer Micelles as Carriers for Anti-Cancer Drugs ................... 400<br />

19.5 Conclusions and Future Perspectives................................................................................ 404<br />

References .................................................................................................................................... 404<br />

19.1 INTRODUCTION<br />

Recently, polymer micelles derived from amphiphilic block copolymers in an aqueous phase have<br />

attracted attention as a promising formulation for poorly water-soluble drugs. 1–4 Block copolymer<br />

micelles are nanosized particles with a typical core-shell structure. 5 The core solubilizes the<br />

hydrophobic drugs, while the corona allows the suspension of micelles in an aqueous medium.<br />

The use of block copolymer micelles as drug-carrying vehicles was proposed <strong>by</strong> Ringsdorf’s group<br />

in the 1980s. 6 The rationale for incorporating low-molecular-weight drugs into micelles is to<br />

overcome the problems of drug formulations such as toxic side effects, poor pharmacokinetics,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

385


386<br />

Nanotechnology for Cancer Therapy<br />

and limited solubility in water. Hydrophobic drugs can be solubilized into hydrophobic core<br />

structures of polymeric micelles, and solubilized at higher concentrations than their intrinsic<br />

water-solubility. 7–10 The chemical composition of polymeric micelles is attractive because it can<br />

be tailored to have desirable physicochemical properties for drug solubilization. Therefore, various<br />

amphiphilic block copolymers have been synthesized and investigated for micelle formulations of<br />

poorly soluble drugs. 11–14 In most polymeric micelles, hydrophobic drugs can be incorporated into<br />

the hydrophobic core of micelles <strong>by</strong> hydrophobic interaction and other additional interactions, such<br />

as the metal–ligand coordination bond, receptor–ligand interaction, and the electrostatic<br />

interaction. 15–17 It is believed that the more compatible drugs are with the cores of the micelles,<br />

the greater their ability to be dissolved.<br />

As drug carriers, polymeric micelles have been widely investigated for a diverse class of anticancer<br />

agents including paclitaxel, adriamycin, and methotrexate. 3,12,18,19 However, several limitations,<br />

for example, limited drug-solubilizing ability, poor stability in water after drug loading, and<br />

the lower stability with the higher loading content, are known to limit successful clinical<br />

applications. 20,21<br />

Hydrotropy has many advantages in enhancing the water solubility of poorly soluble drugs.<br />

Hydrotropes (or hydrotropic agents) self-associate to form noncovalent assemblies of nonpolar<br />

microdomains to solubilize poorly water-soluble solutes. 22 The high concentration of hydrotropes<br />

(greater than 1 M) is a key factor in enhancing water-solubility of poorly soluble solutes. 23,24<br />

Hydrotropes often exhibit a higher selectivity in solubilizing guest hydrophobic molecules than<br />

in surfactant micelles. Thus, identifying hydrotrope structures that effectively solubilize a specific<br />

drug molecule is important. Recently, Park et al. have examined a number of hydrotropes for<br />

solubilization of paclitaxel, a model poorly soluble anti-cancer agent. 25 Through screening the<br />

effective structures of the hydrotropes for paclitaxel solubilization, N,N-diethylnicotinamide<br />

(DENA) and N-picolylnicotinamide (PNA) were found to be the best hydrotropes. They increased<br />

the water-solubility of paclitaxel <strong>by</strong> 3–5 orders of magnitude over its normal solubility in pure<br />

water (about 0.3 mg/mL). They also examined polymers and hydrogels, based on DENA and PNA<br />

hydrotropes, to develop new polymeric solubilizing systems maintaining the benefits of hydrotropy.<br />

26 The hydrotropic property of the hydrotropes was maintained in their polymeric forms, and<br />

the highly localized concentration of the hydrotrope in polymers and hydrogels was found to be a<br />

main contributor to effective solubilization of paclitaxel. However, upon dilution in aqueous media,<br />

drugs solubilized in hydrotropic polymers precipitate due to the low physical stability of the<br />

formulations. Thus, a need exists for the hydrotropic formulations with a high stability.<br />

To overcome the limitations of current polymeric micelles, hydrotropic polymer micelle<br />

systems that have had core components designed with a high solubilizing capacity for poorly<br />

soluble anti-cancer agents, like paclitaxel, show enhanced long-term stability even at high drug<br />

loading. 27 The key in the polymer system designs is to introduce the identified structure of hydrotropes<br />

for a specific drug to the drug-solubilizing micellar cores. The benefit of using selfassembled<br />

structures is the congestion of hydrotropic moieties with high local concentration at<br />

the micellar inner core. To date, many polymeric micelles have shown limited solubilizing capacity<br />

for paclitaxel, and, in most cases, the maximum content of paclitaxel loaded in micelles was around<br />

20 wt%. 20,21 In addition to the limited loading capacity, the stability of the drug-loaded polymeric<br />

micelles is poor. Stability decreases as the drug loading increases. The poor colloidal stability of<br />

existing polymeric micelles is normally caused <strong>by</strong> the enhanced hydrophobicity of micelles after<br />

solubilization of paclitaxel, leading to aggregation of micelles. Because the hydrotropic moiety is<br />

relatively hydrophilic, it retains the hydrophilicity of the colloidal micelles even after loading of<br />

large amounts of paclitaxel.<br />

This chapter gives a general review on recent developments of polymer micelles for cancer<br />

chemotherapeutics, the basic concept of hydrotropic solubilization, and a systemic rationale of<br />

hydrotropic polymer micelle systems. In particular, it focuses on the design strategy and unique<br />

properties of hydrotropic micelle formulations for the delivery of poorly soluble anti-cancer drugs.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 387<br />

19.2 GENERAL REVIEW OF POLYMER MICELLE DRUG CARRIERS<br />

Micelles are defined as colloidal dispersions, including a category of a dispersed system that<br />

consists of the particulate matter or the dispersed phase in a continuous phase or dispersion<br />

medium. 28,29 Micellization phenomena have been studied widely in the field of drug delivery.<br />

Amphiphilic small molecules have been utilized for preparing micelles, but amphiphilic macromolecules<br />

are better than the small molecules because of their lower critical micelle concentration<br />

(CMC) and higher stability. Polymeric micelles represent a separate class of micelles. Polymeric<br />

micelles consist of amphiphilic block copolymers which include multi-block copolymers, AB-type<br />

di-block copolymers, and ABA-type tri-block copolymers with a hydrophobic core (A or B<br />

segment) and a hydrophilic shell (A or B segment). Representative block copolymers are summarized<br />

in Table 19.1. 5,9,12 The amphiphilic block copolymer forms spherical micelles in aqueous<br />

environments with a hydrophilic shell and hydrophobic core parts (Figure 19.1). In general, the core<br />

part represents a molten-liquid globule and a swollen, hydrophilic corona. These micelles have a<br />

high solubilization capacity of poorly soluble drugs, and good potential to minimize drug<br />

degradation, loss, and harmful side effects. In this chapter, polymeric micelles for drug delivery<br />

applications are briefly summarized in terms of methods/mechanism of micelle formation, influencing<br />

factors of stabilization, and solubilization of poorly soluble drugs.<br />

19.2.1 PREPARATIVE METHODS OF POLYMER MICELLES BASED ON BLOCK COPOLYMERS<br />

There are two ways to prepare polymeric micelles: the direct dissolution method and the dialysis<br />

method (Figure 19.2). 1 One can select a method depending on the block copolymer’s solubility in<br />

an aqueous medium. As for direct dissolution, the block copolymer is simply dissolved in water or<br />

buffers at a concentration above its CMC. If needed, the temperature is elevated during the micelle<br />

formation. A good example of this method is poly(ethylene oxide)-block-poly(propylene oxide)block-poly(ethylene<br />

oxide) (PEO–PPO–PEO; Pluronics w ). 30,31 If block copolymers are not easily<br />

water soluble, the dialysis method is useful to prepare polymeric micelles. In this case, block<br />

copolymer is dissolved in a water-miscible organic solvent such as dimethylformamide (DMF),<br />

dimethylsulfoxide (DMSO), and acetonitrile. The mixed solution is then dialyzed against distilled<br />

water. During the dialysis, the organic solvent is removed <strong>by</strong> exchanging the outer medium<br />

(distilled water). The pore size of the semi-permeable membrane (usually SpectraPor w ) should<br />

be carefully selected, based on the expected size of polymeric micelles. The choice of organic<br />

solvent is important because the solvent effects the size and size-population distribution of polymeric<br />

micelles. La et al. reported that the use of DMSO as the organic solvent gave rise to PEO-bpoly(b-benzyl<br />

L-aspartate) (PEO–PBLA) micelles, of which size was only 17 nm. 32 However, only<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

TABLE 19.1<br />

Representative Block Copolymers for Poorly Soluble Drugs<br />

Block Copolymers<br />

Pluronics w (PEO–PPO–PEO)<br />

Methoxy-PEG-b-poly(D,L-lactide) (PEG–PLA)<br />

PEG-b-poly(lactic acid-co-glycolic acid)-b-PEG<br />

Methoxy-PEG-b-polycaprolactone (PEG–PCL)<br />

Poly(b-maleic acid)-b-poly(b-alkylmaleic acid alkyl ester)<br />

Poly(N-isopropyl acrylamide)-b-PEG<br />

Poly(aspartic acid)-b-PEG<br />

PEG-b-phosphatidyl ethanolamine


388<br />

Hydrophilic block<br />

Single polymer chain<br />

Hydrophobic block<br />

Polymer micelle<br />

FIGURE 19.1 General scheme of micelle formation using amphiphilic block copolymers.<br />

6% of the PEO–PBLA formed the micelle. When dimethylacetamide (DMAc) was used as<br />

the organic solvent, the size was 19 nm with a high yield. Therefore, one should examine the<br />

effect of organic solvents on the size and size distribution, as well as the composition, of<br />

block copolymers.<br />

19.2.2 CHARACTERIZATION AND PROPERTIES OF THE POLYMERIC MICELLES<br />

The micellization of di- or tri-block copolymers is usually characterized <strong>by</strong> the following methods:<br />

1. Dye-solubilization methods using 1,6-diphenyl-1,3,5-hexatriene (DPH) and pyrene<br />

(Figure 19.3) 11,33–37<br />

2. Static and dynamic light-scattering measurements. 32,38,39<br />

The dye-solubilization method is useful to determine CMC, local viscosities, and polarity of the<br />

micellar core. 40 For example, DPH is preferentially partitioned into the hydrophobic core of<br />

micelles with the formation of micelles, which causes an increase in the absorbance of the dyes.<br />

When the concentration of the block copolymers changes, CMC can be determined as the crosspoint<br />

of the extrapolation of the change in absorbance over a wide concentration range. Pyrene has<br />

(a)<br />

(b)<br />

Dissolve block copolymer<br />

in aqueous medium<br />

(If needed, increasing temperature)<br />

Dissolve the copolymer<br />

in organic solvents<br />

Put into a dialysis tube<br />

Nanotechnology for Cancer Therapy<br />

Dialysis against water<br />

FIGURE 19.2 Preparation of micelles <strong>by</strong> direct dissolution (a) and dialysis methods (b).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 389<br />

1,6-Diphenyl-1,3,5-Hexatriene (DPH) Pyrene<br />

FIGURE 19.3 Chemical structures of DPH and pyrene.<br />

been well-studied as a fluorescent probe to determine CMC, as well as DPH. 4,41 Below CMC, a<br />

small amount of pyrene is solubilized in water (the saturated concentration is as low as 6!10 K7 M).<br />

In the presence of polymeric micelles, pyrene is preferentially solubilized in the nonpolar micelle<br />

core. By increasing the concentration of the block copolymer in the presence of pyrene, an increase in<br />

emission intensity can be seen, at which point pyrene becomes associated with the micelle core.<br />

When the concentration of the block copolymers changes, CMC can be determined as the point from<br />

which to extrapolate the change in absorbance in a wide range of the concentration.<br />

Weight-average molecular weight (M W) of micelles is calculated <strong>by</strong> the results of static lightscattering<br />

(SLS) measurements using a De<strong>by</strong>e plot 39 :<br />

KðCKCMCÞ=ðRKR CMCÞ Z 1=M W C2A 2ðCKCMCÞ<br />

where K indicates 4p 2 ðdn=dcÞ 2 =N Al 4 , R and R CMC the excess Rayleigh ratios at concentration C and<br />

CMC, n is the refractive index of solution at CMC, dn/dc is the refractive index increment, N A is<br />

Avogadro’s number, l is the wavelength of the laser light, and A2 is the second virial coefficient.<br />

Hydrodynamic radii (RH) of micelles can be determined <strong>by</strong> dynamic light-scattering (DLS)<br />

measurements using the Stokes–Einstein equation 39,41 :<br />

RH Z kBT=ð3phDÞ D Z G=K 2<br />

where k B is the Boltzmann constant, T is the absolute temperature, h is the solvent viscosity, D is the<br />

diffusion coefficient obtained from the average characteristic line width (G), and K 2 is the magnitude<br />

of the scattering vector. Because polymeric micelles are a kind of spherical assembly, the<br />

diffusion coefficients should be independent of the detection angle, due to the undetectable<br />

rotational motions.<br />

19.2.3 KEY PARAMETERS FOR MORPHOLOGY AND STABILIZATION OF THE POLYMERIC MICELLES<br />

As for AB-type di-block copolymers, one can imagine that the di-block copolymer can form<br />

spherical micelles in an aqueous solution. If the hydrophilic block is too long, the di-block copolymers<br />

exist in water as a unimer or a macromolecular micelle. On the other hand, if the<br />

hydrophobic block is too long, it can show nonmicellar morphology, such as rods and lamellar<br />

structures. Eisenberg and his group reported that different morphologies are formed at equilibrium,<br />

near-equilibrium, and nonequilibrium conditions. 42 When the length of the hydrophobic segment is<br />

significantly shorter than that of the hydrophilic part, the block copolymer usually forms nonspherical<br />

micelles. The formation of “crew-cut” aggregates has been suggested <strong>by</strong> a force-balance effect<br />

between the hydrophobic part’s degree of stretching, the interfacial energy of the micelle core with<br />

water, and the interaction between hydrophilic chains as a shell. 42–44 Copolymer compositions,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


390<br />

concentrations, and organic solvents used for the micelle preparation significantly affect the<br />

morphology.<br />

From the pharmaceutical point of view, stability of polymeric micelles in the blood stream is<br />

important because administration of micelles into the body always accompanies dilution, which<br />

causes dissociation of the micelles. 1 The stability of polymeric micelles both in vitro and in vivo<br />

depends on their CMC values. However, the CMC values have been used as a thermodynamic<br />

stability: Equilibrium between unimers and micelles can be shifted below or above the CMC. On<br />

the other hand, kinetic stability, which represents the actual rate of micelle dissociation below the<br />

CMC, depends on the size of the hydrophobic part, the physical state of the micelle core, and<br />

the hydrophilic/hydrophobic ratio. The hydrophobic block plays an especially critical role. For<br />

example, an increase in the length of the hydrophobic block at a given length of the hydrophilic<br />

block causes a significant decrease in the CMC value, and an increase in the stability of the micelle.<br />

On the other hand, an increase in the length of hydrophilic block leads to only a small rise of CMC<br />

value. Additionally, CMC values of di-block copolymers are generally lower than those of tri-block<br />

copolymer of the same molecular weight and hydrophilic/hydrophobic ratio. Thus, the molecular<br />

design of block copolymers for polymeric micelles should be optimized for pharmaceutical applications.<br />

To increase stability, unimolecular micelles that mimic polymeric micelles regarding their<br />

morphological properties have been proposed since 1991. 45,46 These micelles are intrinsically<br />

stable upon dilution, because their formation is independent of polymer concentration. Both<br />

dendrimers and star polymers have been designed as unimolecular micelles. 45,47–49<br />

19.2.4 DRUG LOADING, SOLUBILIZATION, AND DRUG RELEASE<br />

Drug loading methods mostly depend on the methods of micelle formation. 50,51 In the case of<br />

micelles prepared <strong>by</strong> the direct dissolution in water, a stock solution of drugs in some organic<br />

solvents is prepared in an empty vial. Then the organic solvent is allowed to evaporate, followed <strong>by</strong><br />

adding an aliquot of water containing the block copolymer. The oil-in-water emulsion method is<br />

another method: the drug, in a nonpolar solvent such as chloroform, is added dropwise to the water<br />

containing the block copolymer. In the case of the dialysis method, the drug is mixed with the block<br />

copolymer in common organic solvents, and the organic solvents in dialysis bags are exchanged<br />

with a large amount of water to induce assembly of micelles. Upon micellization, the poorly soluble<br />

drugs are trapped in the hydrophobic inner cores of micelles.<br />

It is known that the solubilization of poorly soluble drugs in the micelle core strongly depends<br />

on the types and efficacy of the interactions between the solubilized drug and the hydrophobic<br />

micelle core. 1 Drug-loading capacity depends on the compatibility between the loaded drug and the<br />

hydrophobic core. Drug characteristics such as polarity, hydrophobicity, and charge often affect<br />

compatibility; consequently, the structure and the length of hydrophobic block should be carefully<br />

examined to determine maximum compatibility with any poorly soluble drugs. A good parameter<br />

that has been used to assess compatibility is the Flory–Huggins interaction parameter, csp. This<br />

parameter is described <strong>by</strong> the following equation 1 :<br />

c sp Z ðd sKd pÞ 2 V s=RT;<br />

Nanotechnology for Cancer Therapy<br />

where csp is the interaction parameter between the solubilized drug (s) and the hydrophobic block<br />

part (p), d is the Scatchard–Hildebrand solubility parameter of the hydrophobic block part, and Vs is<br />

the molar volume of the solubilized drug. If the csp shows low value, the compatibility between<br />

drug and hydrophobic block part is great. By using this parameter, Gadelle et al. suggested the<br />

following mechanism for solubilization of aromatic solutes in PEO-b-PPO-b-PEO 28 :<br />

1. The addition of apolar solutes (drugs) promotes aggregation of the block<br />

copolymer molecules.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 391<br />

2. The micelle core contains some water molecules.<br />

3. Solubilization is initially a replacement process in which water molecules are displaced<br />

from the micelle core <strong>by</strong> the drug.<br />

Drug-release kinetics generally depend upon the rate of the drug’s diffusion from the micelles.<br />

There are several factors that affect the release rate: Polymer–drug interactions, localization of the<br />

drug within the micelle, the physical state of the micelle core, the length of the hydrophobic part of<br />

the block copolymer, the molecular volume of the drug, and the physical state of the drug in the<br />

micelle. 1 For instance, if the drug is located in the micelle core as a crystal, it may act as a<br />

reinforcing filter. If the magnitude of interaction between the copolymer and the surface of crystallite<br />

is strong, the drug crystal may cause the glass-transition temperature to increase. Thus,<br />

solubilization of the drug, which depends on the characteristics of both the drug and the block<br />

copolymer, is important in designing polymeric micelles for poorly soluble drugs, such as anticancer<br />

drugs.<br />

19.3 POLYMER MICELLES FOR CANCER CHEMOTHERAPY<br />

19.3.1 POLYMER MICELLES AS CARRIERS OF ANTI-CANCER DRUGS<br />

Significant progress in cancer therapy has been made with the development of new anti-cancer<br />

agents and related technologies. However, the major challenge remains in delivery technologies<br />

that can effectively, selectively deliver anti-cancer agents to tumor sites, and avoiding systemic<br />

toxicity and adverse effects to normal tissues or cells. Recently, polymer micelles have attracted<br />

attention as a novel drug carrier system, in particular for anti-cancer agents, due to its various,<br />

promising properties. These nanoscaled delivery systems have been developed to demonstrate a<br />

series of required properties as drug carriers, such as biocompatibility, stability both in vitro and<br />

in vivo, targeting, and drug loading capacity. Recent research into the development of polymer<br />

micelle delivery system has focused on two issues, drug solubilization and targeted delivery<br />

system, to enhance the efficacy of the delivered drug. 52–55<br />

19.3.2 POLYMER MICELLES FOR SOLUBILIZATION OF POORLY SOLUBLE ANTI-CANCER DRUGS<br />

Because many anti-cancer drugs and drug candidates are water-insoluble or poorly water-soluble,<br />

applicability of solubilizing systems to these drugs is indispensable. For instance, paclitaxel is one<br />

of the most effective anti-cancer agents, and currently formulated with use of surfactants, including<br />

Cremophor EL, due to its extremely poor water solubility. 56 Despite good efficacy of the formulation,<br />

its clinical use is limited <strong>by</strong> serious side effects resulting from use of Cremophor EL. 57–59<br />

Diverse approaches—such as chemical and physical modification, use of a cosolvent, and emulsification—have<br />

been explored for increasing the aqueous solubility of poorly soluble drugs. 60–66<br />

Although several systems have shown high solubilizing effects, they have been limited <strong>by</strong> their<br />

stability and toxicity. Thus, development of nontoxic and stable effective solubilizing systems for<br />

poorly soluble drugs is important for enhancing drug efficacy with high bioavailability.<br />

Over the past decade polymer micelles have been extensively investigated as a potent drug<br />

carrier that can effectively solubilize poorly soluble anti-cancer drugs. Because polymer micelles<br />

have been made with biocompatible polymers such as poly(ethylene oxide), poly(D,L-lactide), etc.,<br />

they are much less toxic than other solubilizing agents such as cosolvents or surfactants. 20,58<br />

Additionally, recent advances in polymer synthesis make it possible to sophisticate the design of<br />

polymer composition and structure so that the resulting micelle shows high solubilizing capacity<br />

with prolonged stability.<br />

Presently a number of polymer micelle systems with different compositions and structures have<br />

been tried to solubilize and deliver various anti-canter drugs. Several examples of the polymer<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


392<br />

micelles used for solubilization of poorly water-soluble anti-cancer drugs are summarized in<br />

Table 19.2. 2–14,19,20,42,52,54 Drug-loading tests have shown that, in most polymer micelle<br />

systems, not only molecular design of block copolymers, but also the drug incorporation method<br />

are crucial factors to increase the solubilizing capacity. 1,14<br />

19.3.3 TARGETING SYSTEMS USING POLYMER MICELLES<br />

Unique core-shell structure of polymer micelle systems with a nanoscaled size may ensure<br />

prolonged circulation times that are favorable for passive drug targeting. 10 Furthermore, polymer<br />

micelles can be selectively accumulated to tumor sites <strong>by</strong> the enhanced permeability and retention<br />

(EPR) effect, which makes them useful as a tumor-targeting system. An alternative, passivetargeting<br />

strategy is to make smart micelles using stimuli-responsive amphiphilic block copolymers<br />

that can sense and respond to the unique tumor environment. 54,67 Kataoka et al. have developed a<br />

novel intracellular pH-sensitive polymeric micelle drug carrier that controls the systemic, local, and<br />

subcellular distributions of pharmacologically-active drugs using amphiphilic block copolymers,<br />

poly(ethylene glycol)–poly(aspartate hydrazone adriamycin). 68 Because the anti-cancer drug,<br />

adriamycin, is conjugated to the hydrophobic segments through acid-sensitive hydrazone links,<br />

this polymer micelle can stably preserve drugs under physiological conditions (pH 7.4), but selectively<br />

release the drug <strong>by</strong> sensing the intracellular pH decrease in endosomes and lysosomes (pH 5–<br />

6).<br />

The active targeting is usually achieved <strong>by</strong> attaching specific-targeting ligand molecules, such<br />

as monoclonal antibodies, to the micelle surface, which provides a preferential accumulation in the<br />

tumor sites or cells. Such ligand and other targeting moieties are introduced in other literatures. 69<br />

When linked with tumor-targeting ligands, polymer micelles can be used to target tumor sites with<br />

TABLE 19.2<br />

Use of Polymer Micelle Systems for Solubilization of Poorly Soluble Anti-Cancer Agents<br />

Polymer Composition a<br />

Nanotechnology for Cancer Therapy<br />

Loaded Drug Loading Content (wt%) b<br />

PEG–PDLLA Paclitaxel 5–25<br />

Methotrexate 3.7–2.8<br />

PEG–PDLLACL Paclitaxel 15–25<br />

PEG–PGACL Paclitaxel 8–11<br />

PEtOz–PCL Paclitaxel 0.5–7.6<br />

PEG–PCL Paclitaxel 4.1–20.8<br />

PEG–pHPMAmDL Paclitaxel 22<br />

PVP-b-PDLLA Paclitaxel 5<br />

Docetaxel 4<br />

Teniposide 10<br />

Etoposide 20<br />

PEG-poly(aspartate) Camptothecin 13<br />

PEG–P(Asp(ADR) Doxorubicin Ellipticine 8.4–7.8<br />

PEG–PBTMC Doxorubicin 10<br />

Poly(NIPAAm-co-DMAAm)-PLGA 2–8<br />

a<br />

Calculated <strong>by</strong>: (weight of loaded drug)!100/(weight of drug-loaded micelle).<br />

b<br />

PDLLA, Poly(D,L-lactide); PDLLACL, Poly(D,L-lactide-co-caprolactone); PGACL, Poly(glycolide-co-caprolactone);<br />

PEtOz, poly(2-ethyl-2-oxazoline); pHPMAmDL, poly(N-(2-hydroxypropyl) methacrylamide lactate); PVP, poly(N-vinylpyrrolidone);<br />

P(Asp(ADR), adriamycin-conjugated poly(aspartic acid) block copolymer; PBTMC,<br />

poly(5-benzyloxytrimethylene carbonate); NIPAAm, N-isopropylacrylamide; DMAAm, N,N-dimethylacrylamide; PLGA,<br />

poly(D,L-lactide-co-glycolide).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 393<br />

high affinity and specificity. Recently a multifunctional polymer micelle system has been proposed<br />

to endow enhanced tumor selectivity and endosome-disruption property on the carrier. 68 The<br />

polymer micelle is designed to expose the cell interacting ligand (biotin) only under slightly<br />

acidic environmental conditions of various solid tumors, and show pH-dependent dissociation,<br />

causing the enhanced release of the loaded drug from the carrier in early endosomal pH.<br />

19.3.4 OTHER APPLICATIONS OF POLYMER MICELLES FOR CANCER THERAPY<br />

Polymer micelles have supramolecular functionalities that are not available from either individual<br />

polymer chains or bulk polymer solids. Based on supramolecular architecture, polymer micelles<br />

have much more surface area and a separate core that can be available for chemically or physically<br />

introducing other active agents, including optical, radioisotopic, magnetic diagnostic, and photosensitive<br />

agents. 70,71 Although the delivery of chemotherapeutic agents <strong>by</strong> polymer micelle<br />

systems is relatively well established, the application for delivering imaging and photosensitive<br />

agents remains in the early stages of research. 47,70,71 The polymer design and targeting strategies<br />

can be also utilized in delivery of other active agents for cancer treatments. Recently, photodynamic<br />

therapy (PDT) has been considered as a promising method of treating cancer and other diseases. 72<br />

PDT involves the systemic administration of photosensitizers followed <strong>by</strong> a local application of<br />

light, introducing photochemical reactions that generate cytotoxical substances that produce<br />

necrosis of a cancer cells. 70 Polymer micelles can be a good candidate for delivery of photosensitizers<br />

in PDT because of their high solubilizing capacity for hydrophobic drugs and potential<br />

targeting property. The delivery of photosensitizers <strong>by</strong> polymer micelles may allow selective<br />

accumulation to solid tumor tissue and resultant higher photocytotoxicity with reduced side<br />

effects. 47,72 Also, hydrophobic photosensitizers that are poorly soluble or tend to aggregate in<br />

aqueous environments can be effectively solubilized in polymer micelle system, showing high<br />

photodynamic efficacy <strong>by</strong> overcoming problems from intrinsic poor solubility. 70<br />

19.4 HYDROTROPIC POLYMER MICELLES<br />

19.4.1 HYDROTROPY AND HYDROTROPIC AGENTS<br />

Hydrotropy is a collective molecular phenomenon describing a solubilization process where<strong>by</strong> the<br />

presence of large amounts of a second solute, called hydrotropes, results in an increased aqueous<br />

solubility of a poorly soluble compound. 22,73 Hydrotropic agents are a diverse class of substances<br />

with wide industrial usage, including solubilizing agents in drug formulations, agents in the separation<br />

of isomeric mixtures, catalysts in heterogeneous phase chemical reactions, and fillers in<br />

cleaning formulations and cosmetics. Some examples of hydrotropic agents are nicotinamide<br />

and its derivatives, anionic benzoate, benzosulfonate, neutral phenol such as catechol and resorcinol,<br />

aliphatic glycolsulfate, and amino acid L-proline. 22,74 Figure 19.4 shows a variety of<br />

compounds that are classified as hydrotropic agents. Hydrotropic agents are characterized <strong>by</strong> a<br />

short, bulky, and compact moiety such as an aromatic ring, whereas surfactants have long hydrocarbon<br />

chains. In general hydrotropic agents have a shorter hydrophobic segment, leading to a<br />

higher water solubility, than that of surfactants. Other examples of hydrotropic materials<br />

are sodium gentisate, sodium glycinate, sodium toluate, sodium naphtoate, sodium ibuprofen,<br />

pheniramine, lysine, tryptophan, isoniazid, and urea. 22,23 At concentrations higher than the<br />

minimal hydrotrope concentration, hydrotropic agents self-associate and form noncovalent<br />

assemblies of lowered polarity (i.e., nonpolar microdomains) to solubilize hydrophobic solutes.<br />

The self-aggregation of hydrotropic agents is different from surfactant self-assemblies (i.e.,<br />

micelles), in that hydrotropes form planar or open-layer structures instead of forming compact<br />

spheroid assemblies.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


394<br />

N<br />

O<br />

NH2<br />

Nicotinamide Sodium nicotinate<br />

Sodium benzoate<br />

Sodium salicylate<br />

CH3<br />

CH3<br />

O<br />

OH<br />

O - Na +<br />

SO3 - Na +<br />

Sodium xylenesulfonate<br />

H2N<br />

19.4.2 HYDROTROPIC AGENTS IN PHARMACEUTICS<br />

N<br />

O<br />

OH<br />

O - Na +<br />

O<br />

O - Na +<br />

Sodium p-aminosalicylate<br />

Resorcinol<br />

Delivery of poorly water-soluble drugs <strong>by</strong> oral route has been limited <strong>by</strong> poor absorption from the<br />

gastrointestinal tract (GI). 75 One of the governing factors in the transport of drugs across the<br />

intestinal barrier is the carrier-mediated efflux mechanism, which is referred to as multidrug<br />

resistance (MDR). This phenomenon is mediated <strong>by</strong> the increased expression of energy-dependent<br />

drug efflux proteins such as P-glycoproteins, multidrug-resistance-associated proteins, and lungresistance<br />

proteins. Since the excretion of the absorbed drugs <strong>by</strong> this mechanism is the saturable<br />

process, the solubility enhancement of poorly water-soluble drugs might overcome the barrier. At<br />

high concentration of drugs, the efflux protein will be saturated, there<strong>by</strong> resulting in the increase of<br />

the absorbed amount of drugs and bioavailability. The hydrotrope approach is a promising new<br />

method with great potential for poorly soluble drugs. Using hydrotropic agents is one of the easiest<br />

ways of increasing water-solubility of poorly soluble drugs because it only requires mixing the<br />

drugs with hydrotropes in water. 24,25 Hydrotropic agents are commonly used to enhance the watersolubility<br />

of poorly soluble drugs, and in many instances the water-solubility of these drugs is<br />

increased <strong>by</strong> orders of magnitude. The use of hydrotropic agents offers numerous benefits over<br />

other solubilization methods such as micellar solubilization, cosolvency, and salting-in. 74 To date,<br />

various hydrotropic agents have been utilized to enhance the aqueous solubility of many hydrophobic<br />

drugs including paclitaxel, ketoprofen, diazepam, allopurinol, indomethacin, and<br />

griseofulvin. In most cases, the high concentration of hydrotropes is key to increasing water<br />

solubility of poorly soluble drugs. Thus, the approach to increase the local concentration of hydrotropes<br />

may provide an opportunity to develop new hydrotrope-based solubilizing systems. Despite<br />

the advantages of hydrotropes, application of low molecular weight hydrotropes in drug delivery<br />

has not been practical because it may result in absorption of a significant amount of hydrotropes<br />

themselves into the body, along with the drug. Recently polymeric forms of hydrotropes, such as<br />

OH<br />

CH3<br />

FIGURE 19.4 Chemical structures of hydrotropes used in the literature.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

O<br />

O - Na +<br />

SO3 - Na +<br />

Sodium p-toluenesulfonate<br />

H 3C (CH 2) 3 OCH 2CH 2 OSO 3 - Na +<br />

Sodium butyl monoglycol sulfate


Hydrotropic Polymer Micelles for Cancer Therapeutics 395<br />

hydrotropic polymers and hydrogels, were developed to overcome the drug formulations based on<br />

low molecular weight hydrotropes. 26 The polymeric hydrotropes were found to maintain the hydrotropic<br />

activity of its low molecular weight hydrotropes. Their solubilizing ability for paclitaxel was<br />

well examined, and the solubilization mechanism correlated with the assembly of polymeric hydrotropes.<br />

However, the low physical stability of drug formulations is also reported to limit the<br />

practical applications of these formulations.<br />

19.4.3 MECHANISTIC STUDIES AND STRUCTURE–PROPERTY RELATIONSHIP OF HYDROTROPIC<br />

SOLUBILIZATION<br />

The term hydrotropy does not mean a specific mechanism, but implies a collective solubilization<br />

phenomenon that is incompletely understood. There have been various theoretical and experimental<br />

studies aimed at explaining hydrotropic solubilization. Most of proposed mechanisms<br />

can be classified into following two schemes. The first one involves the complex formation<br />

between the hydrotrope and the solute. As a representative example, nicotinamide—a nontoxic<br />

vitamin B 3—has been shown to enhance the solubilities of a wide variety of hydrophobic drugs<br />

through complexation. Using nicotinamide and its derivatives, such as N-methylnicotinamide and<br />

N,N-diethylnicotinamide (DENA), Rasool et al. described the aromaticity of pyridine ring, which<br />

might promote the stacking of molecules through its planarity, as a most significant contributor in<br />

complexation, because the abilities of aromatic amide ligands to enhance the aqueous solubilities of<br />

tested drugs was higher than those of the aliphatic amide ligands. 76 The other mechanism for<br />

hydrotropic solubilization is self-association of the hydrotrope in an aqueous phase. This view is<br />

supported <strong>by</strong> experimental data, proving that some hydrotropes, including nicotinamide and<br />

aromatic sulfonates, associate in aqueous solutions. Using nicotinamide-riboflavin system,<br />

Kildsig et al. showed that the self-association of nicotinamide contributed to the solubility increase<br />

of riboflavin, rather than complexation between two species. 77<br />

Each hydrotropic agent is effective in increasing the water solubility of selected hydrophobic<br />

drugs, and no hydrotropic agents were found that are universally effective. Therefore, the structure–<br />

activity relationship between selected pairs of the hydrotropic agent and the drug is another<br />

important concern to discuss. Park et al. reported on the intensive studies of the elucidation of<br />

structure–activity relationship for the hydrotropic solubilization of paclitaxel using not only nicotinamide<br />

and its analogues as aromatic amides, but also various ureas as aliphatic amides. 25 They<br />

showed that nicotinamide enhanced the aqueous solubility of paclitaxel to a greater extent than the<br />

aliphatic analogues of nicotinamide such as nipecotamide and N,N-dimethylacetamide. In addition,<br />

the aqueous solubility of paclitaxel was found to be strongly dependent on the alkyl substituent on<br />

the amide nitrogen of nicotinamide. DENA of 3.5 M enhanced the aqueous solubility of paclitaxel<br />

up to 39.07 mg/mL, whereas N,N-dimethylnicotinamide showed paclitaxel solubility of<br />

1.77 mg/mL at the same concentration. They synthesized new hydrotropic agents based on the<br />

structure of nicotinamide <strong>by</strong> varying the substituent to the amide nitrogen with a pyridine ring or an<br />

ally group. The aqueous solution of N-picolylnicotinamide (PNA) (3.5 M), having another pyridine<br />

ring as a substituent on the amide nitrogen, was highly effective in the enhancement of paclitaxel<br />

solubility (29.44 mg/mL), compared with the hydrotropic effect of nicotinamide on the<br />

paclitaxel (0.69 mg/mL). Besides, N-allylnicotinamide highly contributed to the enhancement of<br />

paclitaxel solubility (14.18 mg/mL) compared to that of N,N-dimethylnicotianmide (1.77 mg/mL).<br />

The most interesting find is that the threshold concentration where the association occurred is<br />

consistent with the threshold concentration where paclitaxel solubility began to increase.<br />

19.4.4 DESIGN STRATEGY OF HYDROTROPIC POLYMER MICELLE SYSTEMS<br />

Hydrotropic polymer micelles were firstly developed based on the self-assemblies of amphiphilic<br />

block copolymers containing the hydrophilic block, and the hydrophobic block containing pendent<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


396<br />

hydrotropic moieties. The main strategy for hydrotropic systems is to maximize the local concentration<br />

of hydrotropic structure because the drug solubilizing capacity increases when increasing<br />

the local concentration of hydrotropes. 26 The hydrophobic block possesses hydrotropes assembled<br />

in the limited volume in micellar core region. Considering several hundreds of aggregation numbers<br />

of micelles, the local concentration of hydrotropes would be maximized as compared to other<br />

polymeric hydrotropes. In addition to the enhanced solubilizing ability for poorly soluble drugs,<br />

the intrinsic hydrophilicity of hydrotropic structure can retain the colloidal stability <strong>by</strong> compensation<br />

of enhanced hydrophobicity of drug-loaded polymer micelles. Using atom-transfer radical<br />

polymerization (ATRP) of modified hydrotropes, Park et al. started the synthesis of amphiphilic<br />

block copolymers consisting of PEG and poly(2-(4-(vinylbenzyloxyl)-N,N-diethylnicotinamide)<br />

(PDENA) or poly(2-(4-(vinylbenzyloxyl)-N-picolylnicotinamide) (P(2-VBOPNA)) as shown in<br />

Figure 19.5. 27<br />

PEG-b-PDENA copolymers associated to form polymeric micelles with the hydrotrope-rich<br />

core and the PEG outer shell in aqueous media. The mean hydrodynamic diameters were in the<br />

range of 30–100 nm. In these polymer systems, the solubilization of paclitaxel is based on a synergistic<br />

effect of the unique micelle characteristics and hydrotropic activity. Hydrotropic micelles<br />

demonstrated not only higher loading capacity (up to 37 wt% of paclitaxel) but also enhanced<br />

physical stability in aqueous media. 27 The enhanced stability is due to the intrinsic hydrophilicity<br />

of hydrotropic moieties in the micellar core. The drug loading into the hydrotropic polymeric micelle<br />

core is mainly based on the attractive interactions between the hydrotropic moiety and paclitaxel.<br />

The structure of hydrotropic polymer micelles can be varied <strong>by</strong> introducing a diverse class of<br />

hydrotropic structures into the block copolymers through ATRP. 28,29,78–80 Table 19.3 lists some<br />

of the block and graft copolymers consisting of poly(ethylene glycol) (PEG) and the hydrotropic<br />

polymers that have been synthesized based on the molecular structures of identified hydrotropic<br />

agents for paclitaxel, such as DENA and PNA.<br />

CH3 OCH2CH2 n OH + Br<br />

O<br />

C<br />

N<br />

N<br />

O<br />

O<br />

Cu(I)Br, PMDETA, 80 °C<br />

O<br />

O<br />

N<br />

NH N<br />

CH3 CH Br<br />

TEA/DMAP<br />

CH 2Cl 2<br />

PEG−PDENA<br />

O<br />

O<br />

PEG−P(2-VBOPNA)<br />

Nanotechnology for Cancer Therapy<br />

O<br />

CH3 OCH2CH2 O C<br />

n<br />

CH 3<br />

CH3 OCH2CH2 O C CH CH2 CH Br<br />

n m<br />

CH3<br />

CH3 OCH2CH2 O C CH CH<br />

n 2 CH Br<br />

m<br />

N<br />

N<br />

O<br />

O<br />

O<br />

O<br />

N<br />

CH 3<br />

CH<br />

NH N<br />

FIGURE 19.5 Synthetic routes for PEG–PDENA and PEG–P(2-VBOPNA) block copolymers.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Br


Hydrotropic Polymer Micelles for Cancer Therapeutics 397<br />

The main components necessary for the synthesis of the block copolymers are PEG modified<br />

with bromine or chlorine, and the modified hydrotropic agent with polymerizable vinyl, acryl, and<br />

styryl groups. Table 19.4 lists the useful hydrotropic agents modified with double bond functionality.<br />

The relevant combination of monomers listed in Table 19.4 can lead to a diverse class of the<br />

copolymers capable of making the hydrotropic micelles.<br />

In addition to the structural variation, a new property such as stimuli-responsibility can be<br />

endowed with the hydrotropic polymer micelles. One example is stimuli-responsive hydrotropic<br />

polymer micelles which were derived from PEG–P(2-VBOPNA) copolymers. In this system, PEG<br />

is freely water-soluble, independent of pH, whereas P(2-VBOPNA) is known to be soluble in water<br />

only when the hydrotropic PNA groups are protonated. P(2-VBOPNA) is soluble in low pH but<br />

loses its water-solubility above a critical pH (pH 2.5). Thus, the PEG–P(2-VBOPNA) copolymers,<br />

as expected, dissolved molecularly as unimers at low pH but become amphiphilic above pH 2.5.<br />

The combination of PEG and P(2-VBOPNA) in a block copolymer provides an opportunity to<br />

undergo pH-responsive micellization <strong>by</strong> controlling solution pH. Figure 19.6 shows pH-dependency<br />

of scattering intensity of PEG–P(2-VBOPNA) <strong>by</strong> dynamic light scattering.<br />

As solution pH increases, scattering intensity increases abruptly above pH 2.0, accompanied <strong>by</strong><br />

the formation of micelles around pH 3.0 of which sizes were about 35 nm. As pH of the aqueous<br />

solution of the block copolymers is increased from 2 to 3, the degree of protonation of PNA groups<br />

decreased, and thus the P(2-VBOPNA) block becomes progressively more hydrophobic, as<br />

expected. 1 H NMR analysis is also a good tool to confirm the pH-responsive micellization of<br />

PEG–P(2-VBOPNA) (Figure 19.7).<br />

At pH 1, the block copolymers are fully solvated, and thus the signals assigned for the each<br />

block are clearly visible. On the other hand, at pH 7.4, the resonance peaks of P(2-VBOPNA)<br />

disappear while the signals from PEG are still evident, which indicates the poor solvation and/or<br />

limited molecular motion of P(2-VBOPNA) blocks. This result strongly supports the formation of<br />

micelles consisting of P(2-VBOPNA) as a hydrophobic inner core and PEG as a hydrated outer<br />

shell. Using this unique pH-sensitive micelle system, it may be suggested that PEG-b-<br />

P(2-VBOPNA) allows a unique drug-loading method into polymeric micelles, where drug<br />

molecules are initially solubilized exclusively <strong>by</strong> the interaction with hydrotropic<br />

P(2-VBOPNA) block at low pH, and then drugs are possibly encapsulated into inner core of<br />

micelles upon pH-responsive micellization. The use of hydrotropic block copolymers showing<br />

pH-responsive micellization does not need organic solvents during the drug loading process.<br />

The ideal oral delivery systems for anti-cancer drugs need the two following physicochemical<br />

properties to maximize oral bioavailability of drugs: the fast drug release during the transit of GI,<br />

and the prolonged retention time in the GI tract. The former was known to be the characteristic of<br />

hydrotropic polymer micelles. Thus, if the design of hydrotropic polymer micelles showing<br />

mucoadhesion property is possible, the idealized delivery systems can be developed based on<br />

hydrotropic polymer micelles. The study on the degradation kinetics of paclitaxel in aqueous<br />

TABLE 19.3<br />

Exemplary Copolymers of PEG and Hydrotropic Polymers Synthesized<br />

from Modified Hydrotropic Agents<br />

Poly(ethylene glycol)-block-poly(2-(4-vinylbenzyloxy)-N,N-diethylnicotinamide)<br />

Poly(ethylene glycol)-block-poly(2-(4-vinylbenzyloxy)-N-picolylnicotinamide)<br />

Poly(ethylene glycol)-block-poly(2-(4-vinylbenzyloxy)-nicotinamide)<br />

Poly(oligoethylene glycol methacrylate-co-poly(2-(4-vinylbenzyloxy)-N,N-diethylnicotinamide)<br />

Poly(oligoethylene glycol methacrylate-co-poly(2-(4-vinylbenzyloxy)-N-picolylnicotinamide)<br />

Poly(oligoethylene glycol methacrylate-co-poly(2-(4-vinylbenzyloxy)-nicotinamide)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


TABLE 19.4<br />

Modified Hydrotropic Monomers That Can Be Used in the Synthesis of Hydrotropic Polymer Micelles<br />

N<br />

CH 2 CH<br />

O O<br />

N<br />

HN<br />

O<br />

N<br />

CH2 CH<br />

O<br />

N<br />

HN<br />

3<br />

O<br />

O<br />

O<br />

6-acryloyl-N-picolylnicotinamide 6-(2-(acryloyl) ethoxyethoxyethoxy)-N-picolylnicotinamide<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

N<br />

CH 2 CH<br />

N<br />

HN<br />

O<br />

O<br />

6-(4-vinylbenzyl oxy)-N-picolyl<br />

nicotinamide<br />

CH 2 CH<br />

O O<br />

N<br />

O<br />

N<br />

2-(2-(acryloyl) ethoxyethoxyethoxy)-N, Ndiethylnicotinamide<br />

CH 2<br />

O<br />

N<br />

CH<br />

3 O<br />

O<br />

O<br />

N<br />

2-(4-vinylbenzyloxy)-N, N-diethyl<br />

nicotinamide<br />

CH 2 CH<br />

N<br />

O<br />

O<br />

N<br />

2-acryloyl-N, N-diethylnicotinamide<br />

398<br />

Nanotechnology for Cancer Therapy


Hydrotropic Polymer Micelles for Cancer Therapeutics 399<br />

Scattering intensity (Kcps)<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

0<br />

1<br />

2<br />

3<br />

FIGURE 19.6 Variation of scattering intensity (C) as a function of pH for PEG 5000-P(2-VBOPNA) 2070 at<br />

5 mg/mL (nZ3).<br />

10 9<br />

(a)<br />

pH 1.0<br />

4<br />

pH<br />

d (PPM)<br />

10 9 8 7 6 5 4 3 2 1 0<br />

(b)<br />

8<br />

pH 7.4<br />

7<br />

6<br />

DOH<br />

5<br />

DOH<br />

d (PPM)<br />

FIGURE 19.7 1 H NMR spectra of PEG5000-P(2-VBOPNA)2070 at 20 mg/mL in D2O: (a) pH 1.0; (b) pH 7.4.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

4<br />

5<br />

3<br />

6<br />

2<br />

7<br />

1<br />

8<br />

0


400<br />

Nanotechnology for Cancer Therapy<br />

solutions at different pH values shows that a maximum stability of paclitaxel occurred in the pH 3–5<br />

region. 61 This provides a perfect opportunity to prepare mucoadhesive hydrotropic polymer micelle<br />

formulation for oral paclitaxel delivery. One of the best mucoadhesive polymers is<br />

poly(acrylic acid) (PAA), which is highly adhesion to biological mucosa at a pH lower than 5.<br />

PAA at neutral pH is highly charged and highly water-soluble, and can provide a hydrophilic<br />

segment of the block copolymers. Upon contact with the stomach’s low pH, the polymer<br />

becomes highly mucoadhesive. Alternatively, PAA can be treated at pH 5 to provide the mucoadhesive<br />

property. Because of the highly hydrophilic properties of PAA, it can replace the PEO<br />

segment of the block copolymers. Thus, the design of the block copolymer consisting of PAA<br />

and the hydrotropic block may provide the useful oral formulations of poorly soluble anticancer<br />

agents.<br />

19.4.5 HYDROTROPIC POLYMER MICELLES AS CARRIERS FOR ANTI-CANCER DRUGS<br />

Hydrotropic micelles were first examined for their ability to solubilize paclitaxel, a good example<br />

of an anti-cancer drug with extremely low water solubility. 27 Due to the high potential of paclitaxel,<br />

many formulations have been developed to increase water solubility. Although the polymeric<br />

micelles based on the amphiphilic block copolymers have shown high potential as drug solubilizing<br />

systems, most polymeric micelles have shown limited solubilizing capacity for paclitaxel, and, in<br />

most cases, maximum contents of paclitaxel loaded into micelles was around 20 wt%. 9,20,25<br />

Besides, a simple polymer design may not predict whether the resulting polymer micelles show<br />

high solubilizing capacity. A more serious limitation is the poor stability of paclitaxel-solubilized<br />

polymeric micelles in water, which becomes lower as the content of paclitaxel increases. 20 Of the<br />

properties of solubilizing systems, a high solubilizing capacity and a good physical stability are the<br />

two most important factors in determining whether the drug delivery systems are clinically useful.<br />

The goal of using hydrotropic polymer micelles is to develop a hydrotropic polymeric micelle that<br />

has a high solubilizing capacity for poorly water-soluble drugs, as well as a good long-term<br />

stability. Because an identified hydrotrope for a specific drug is introduced as a core component<br />

in a highly localized way, paclitaxel solubilization may be presented <strong>by</strong> the synergistic effect of<br />

both hydrotropic and micellar solubilization. The poor colloidal stability of existing micelles is<br />

normally caused <strong>by</strong> the enhanced hydrophobicity of micelles after solubilization of paclitaxel.<br />

Hydrotropic moieties, characterized <strong>by</strong> a strong hydrophilic nature, are expected to allow good<br />

stability for paclitaxel-loaded micelles in water. Huh et al. described the superiority of the hydrotropic<br />

polymer micelles relative to current polymeric micelles, and how the hydrotropic polymer<br />

micelles are different from existing systems. 27 These questions can be answered <strong>by</strong> comparing their<br />

drug loading and physical stability. The loading capacity of the normal polymer micelles for poorly<br />

soluble drugs is decided <strong>by</strong> various factors, such as the length of the core-forming polymer, and<br />

compatibility between drugs and the core-forming polymers. 1 Of these factors, the compatibility is<br />

the most significant in determining the solubilizing capacity. One parameter that has been used to<br />

assess the compatibility between solubilizates and the polymer is Flory–Huggins interaction parameter<br />

as described in Section 19.2.4. 1 This value is dependent on both the selected drug and the<br />

polymer. Due to the uniqueness of each drug, no single core-forming block can maximize the<br />

solubilization level for all drugs. Therefore, the first priority is to find or synthesize the right<br />

structure for effective solubilization of selected drugs. However, the number of biocompatible<br />

polymers is limited, and an added difficulty of the synthesis step is the screening of a large<br />

number of the polymer structures for effective solubilization of the selected drug. On the other<br />

hand, the hydrotropic approach is simpler and less labor-intensive, even though a screening process<br />

is also required. Many hydrotropes can be identified for a selected drug <strong>by</strong> a simple mixing<br />

procedure, and the broad range of the chemical structure can be readily screened. 25 The key<br />

concept of the hydrotropic polymer micelles is based on the hydrotrope containing core-forming<br />

polymers. Thus, the systematic design affords more efficient systems for solubilizing poorly soluble<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 401<br />

anti-cancer drugs. Another advantage of this approach is the enhanced physical stability of the<br />

formulations. This property is one measure that distinguishes hydrotropic polymer micelles from<br />

normal polymer micelles. Conventional polymer micelles have a polymer with a strong hydrophobic<br />

nature, specifically the drug solubilization has been expected only from the hydrophobic<br />

interaction between drugs and the inner core. Often, polymer micelles with drugs cannot overcome<br />

the enhanced hydrophobicity and the secondary aggregation between micelles, resulting in precipitation.<br />

From this point of view, the hydrotropic polymer micelles provide the formulations with<br />

good stability, even at a high loading of poorly soluble anti-cancer drugs, due to the hydrophilic<br />

nature of the hydrotropes residing in the micellar core domains. Of course, the same approach can<br />

be used for solubilization of other poorly soluble drugs. Huh et al. reported results on the solubilizing<br />

(loading) effect of the hydrotropic polymer micelle for paclitaxel <strong>by</strong> a dialysis method.<br />

The results are illustrated in Figure 19.8.<br />

The hydrotropic polymeric micelles solubilized paclitaxel at a level of 18.4–37.4 wt%,<br />

depending on the organic solvents used in the dialysis and the initial feed weight ratio of paclitaxel<br />

to the block copolymer. The loading content normally increases to certain feed-weight ratios of<br />

paclitaxel to the block copolymer. However, as the amount of paclitaxel was further increased,<br />

precipitates of unloaded paclitaxel formed during dialysis, resulting in the decreased loading<br />

contents. The maximum loading was observed with initial feed weight ratio of 1:0.5. When acetonitrile<br />

and the feed weight ratio of 1:0.5 were used the loading content was as high as 37.4 wt%,<br />

which was not possible with existing polymeric micelle systems. The maximum loading content of<br />

paclitaxel in a control micelle of PEG 2000–PDLLA 2000 was estimated to be 27.6 wt%, which is close<br />

to the literature value. Loading capacity of PEG–PDENA micelles for paclitaxel was enhanced <strong>by</strong><br />

increasing the block length of PDENA.<br />

Poor physical stability of drug-loaded micelles, which causes serious problems in drug formulation,<br />

has been often addressed. There is often a compromise between loading content and stability<br />

in polymeric micelle systems. Drug-loaded micelles may break up more easily in aqueous media,<br />

resulting in drug precipitation. The physical stability of polymer micelles was investigated <strong>by</strong><br />

several methods and compared. PEG–PLA micelles that have been extensively used to solubilize<br />

paclitaxel demonstrated a high loading capacity, but showed drug precipitation after 24 h.<br />

The initial transparent micelle solution became translucent, or turbid, after 24 h. Similar results<br />

have been reported in the literature. 20,21,25 On the other hand, PEG–PDENA micelle solutions were<br />

observed to be stable for several weeks without drug precipitation at the same or higher loading<br />

contents of paclitaxel.<br />

Paclitaxel content (wt%)<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Acetonitrile DMAc DMF<br />

1 : 0.25<br />

1 : 0.3<br />

1 : 0.35<br />

1 : 0.4<br />

1 : 0.5<br />

1 : 0.6<br />

Control<br />

FIGURE 19.8 Paclitaxel lading contents in hydrotropic polymer micelles and comparison with<br />

PEG2000–PDLLA2000 micelles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


402<br />

The loading content of paclitaxel in polymer micelles was reported as a function of time.<br />

As shown in Figure 19.9, the paclitaxel content in PEG–PDENA micelles was maintained for<br />

more than 30 days. 27<br />

In contrast, the PEG–PLA micelle showed a dramatic decrease in paclitaxel content after three<br />

days, which resulted from drug precipitation. The stability of drug-loaded PEG–PDENA micelles<br />

was further confirmed <strong>by</strong> observing no change in micelle size and scattering intensity using<br />

dynamic light-scattering measurements.<br />

The paclitaxel release profiles of PEG–PDENA and PEG–PLA micelles were compared as<br />

shown in Figure 19.10. 27<br />

For PEG–PDENA micelles, the micelles of 25.9 wt% loading released the most drug within<br />

48 h, whereas the micelles of 31.3 wt% drug loading showed a much faster drug release that was<br />

complete within 24 h. It was found that the faster release was from the micelles with higher drug<br />

loading. The micelles with higher drug loading lead to a relatively lower polymer concentration.<br />

Thus, there is less polymer–drug interaction in the micelles, there<strong>by</strong> resulting in faster release<br />

kinetics. The in vitro release data showed that paclitaxel was released faster from the hydrotropic<br />

polymer micelles than from PEG to PLA micelles. The PEG–PLA micelles required almost 72 h to<br />

release all the loaded paclitaxel. As indicated <strong>by</strong> the higher CMC values of PEG–PDENA micelles,<br />

the hydrotropic polymeric micelles have less hydrophobic cores that may release the loaded<br />

paclitaxel faster than more hydrophobic PLA cores. On the other hand, the release from paclitaxel<br />

bulk powder was negligible, even after 74 h. (Figure 19.10).<br />

The cytotoxicity of anti-cancer agents loaded in polymer micelle formulations is one measure<br />

for determining whether the formulation is effective in delivering the drugs into the tumor cells.<br />

The anti-tumor cytotoxicities of hydrotropic polymeric micelles and other control micelle systems<br />

on various cell lines, as measured <strong>by</strong> ED50, are shown in Table 19.5. The resulting cytotoxicity of<br />

hydrotropic polymer micelles and control micelles clearly show the superior cytotoxic properties<br />

of hydrotropic micelles. The ED50 values of hydrotropic micelles are much lower than PEG–PLA<br />

and PEG–PPA micelles. The most widely used polymeric micelles is PEG–PLA micelles, with the<br />

maximum paclitaxel loading capacity of 24 wt%. Although the hydrotropic polymer micelles have<br />

higher paclitaxel, loading up to 37 wt%, hydrotropic polymer micelles with only 20 and 25 wt%<br />

paclitaxel loading were used to compare their efficacy with that of PEG–PLA micelles.<br />

At equivalent paclitaxel loading, i.e., 25 wt% loading, hydrotropic polymer micelles were<br />

Paclitaxel concentration (%)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

14571g9 eps<br />

PEG−PDENA (25.9 wt% paclitaxel)<br />

PEG−PDENA (18.4 wt% paclitaxel)<br />

PEG−PLA (27.6 wt% paclitaxel)<br />

PEG−PLA (17.6 wt% paclitaxel)<br />

0<br />

0 5 10 15<br />

Time (day)<br />

20 25 30<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 19.9 Changes in paclitaxel concentration of polymer micelles in distilled water. (From Huh, K.M.,<br />

Lee, S.C., Cho, Y.W., Lee, J., Jeong, J.H., and Park, K., Journal of Controlled Release, 101, 59–68, 2005. With<br />

permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 403<br />

0.12<br />

0.08<br />

0.04<br />

0.00<br />

0 20 40 60 80 100<br />

Time (hour)<br />

FIGURE 19.10 Release kinetics from polymer micelles and paclitaxel powders in 0.8 M sodium salicylate at<br />

378C. The total amount of loaded paclitaxel was 0.1 mg. (From Huh, K. M., Lee, S. C., Cho, Y. W., Lee, J.,<br />

Jeong and Park, K., Journal of Controlled Release, 101, 59–68, 2005. With permission.)<br />

substantially more effective. In the case of the MDA231 cell line, hydrotropic polymer micelles<br />

were more than two orders of magnitude more effective. The data clearly indicate that hydrotropic<br />

polymer micelles are not only more stable in aqueous solution, but also more effective for paclitaxel<br />

delivery to the cells.<br />

As mentioned in Section 19.4.2, the enhanced drug solubility can improve oral bioavailability<br />

of drugs <strong>by</strong> suppressing the MDR effect. The excellent solubilizing ability of hydrotropic polymer<br />

micelles for poorly soluble drugs can overcome this barrier, there<strong>by</strong> enhancing drug absorption in<br />

the GI tract. As a preliminary study, PEG–PDENA micelles were examined for the bioavailability<br />

of loaded paclitaxel using an in vivo, chronically-catheterized rat model. Our initial study showed<br />

that oral delivery of paclitaxel using PEG–PDENA micelles could result in the blood paclitaxel<br />

TABLE 19.5<br />

ED 50 (mg/mL) of Paclitaxel and Paclitaxel (PTX)-Loaded Polymeric Micelles on Various<br />

Tumor Cell Lines<br />

Samples a<br />

HT-29<br />

Cancer Cell Lines<br />

MDA231 MCF-7 SKOV-3<br />

Doxorubicin (positive<br />

control)<br />

0.044 0.050 0.773 0.611<br />

Paclitaxel 0.003 0.033 0.043 0.006<br />

PTX-loaded HTM<br />

(25 wt% loading)<br />

0.005 0.002 0.002 0.001<br />

PTX-loaded HTM<br />

(20 wt% loading)<br />

0.006 0.004 ! 0.001 0.008<br />

PTX-loaded PLA-PEG<br />

(24 wt% loading)<br />

0.014 0.305 ! 0.001 0.015<br />

PTX-loaded PEG–PPA 0.012 1.077 0.002 1.543<br />

HTM alone 4.221 5.650 5.767 0.048<br />

PEG–PLA alone 4.672 8.435 5.533 —<br />

PEG–PPA alone 0.277 4.881 6.194 —<br />

a<br />

PTX, paclitaxel; HTM, hydrotropic micelle; PLA, poly(lactic acid); PEG, poly(ethylene glycol); PPA, poly(phenylala-<br />

nine).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


404<br />

concentrations that are clinically significant. To increase oral bioavailability, hydrotropic polymer<br />

micelles with other interesting functions can be developed. The systematic design of hydrotropic<br />

polymer micelles for its faster drug releasing properties, as well as the prolonged retention during<br />

through the GI tract, can provide an opportunity to develop ideal oral delivery systems for poorly<br />

soluble anti-cancer drugs.<br />

19.5 CONCLUSIONS AND FUTURE PERSPECTIVES<br />

Hydrotropic polymer micelle formulation presents an alternative and promising approach in formulation<br />

of poorly soluble drugs. Based on synergistic effect of the micellar characteristics and<br />

hydrotropic activity, the hydrotropic polymer micelles exhibit a high drug loading capacity with<br />

enhanced long-term stability in aqueous media. These unique properties make it highly attractive<br />

for applications of controlled delivery of poorly soluble anti-cancer drugs. The hydrotropic polymer<br />

micelles can be applied to many poorly soluble drugs and drug candidates <strong>by</strong> introducing the<br />

identified hydrotropic structures for specific drug molecules. The hydrotropic polymer micelle is<br />

in the early stages of the development, but due to its unique properties, the systematic design may<br />

generate effective oral formulations applicable to many poorly soluble anti-cancer agents. The high<br />

versatility of the hydrotropic polymer micelle in creating a broad range of drug formulations gives it<br />

the potential for a broad range of formulations in pharmaceutical and biomedical applications.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Allen, C., Maysinger, D., and Eisenberg, A., Nanoengineering blocks copolymer aggregates for drug<br />

delivery, Colloids and Surfaces B: Biointerfaces, 16, 3–27, 1999.<br />

2. Lee, S. C., Kim, C., Kwon, I. C., Chung, H., and Jeong, S. Y., Polymeric micelles of poly(2-ethyl-2oxazoline)-block-poly(3-caprolactone)<br />

copolymer as a carrier for paclitaxel, Journal of Controlled<br />

Release, 89, 437–446, 2003.<br />

3. Yokoyama, M., Okano, T., Sakurai, Y., Fukushima, S., Okamoto, K., and Kataoka, K., Selective<br />

delivery of adriamycin to a solid tumor using a polymeric micelle carrier system, Journal of Drug<br />

Targeting, 7, 171–186, 1999.<br />

4. Lee, S. C., Chang, Y., Yoon, J.-S., Kim, C., Kwon, I. C., Kim, Y.-H., and Jeong, S. Y., Synthesis and<br />

micellar characterization of amphiphilic diblock copolymers based on poly(2-ethyl-2-oxazoline) and<br />

aliphatic polyesters, Macromolecules, 32, 1847–1852, 1999.<br />

5. Kwon, G., Naito, M., Yokoyama, M., Okano, T., Sakurai, Y., and Kataoka, K., Micelles based on AB<br />

block copolymers of poly(ethylene oxide) and poly(beta-benzyl L-aspartate), Langmuir, 9, 945–949,<br />

1993.<br />

6. Bader, H., Ringsdorf, H., and Schmidt, B., Watersoluble polymers in medicine, Die Angewandte<br />

Makromolekulare Chemie, 123/124, 457–485, 1984.<br />

7. Jeong, Y.-I., Cheon, J.-B., Kim, S.-H., Nah, J.-W., Lee, Y.-M., Sung, Y.-K., Akaike, T., and Cho,<br />

C. S., Clonazepam release from core-shell type nanoparticles in vitro, Journal of Controlled Release,<br />

51, 169–178, 1998.<br />

8. Kim, I. S. and Kim, S. H., Development of a polymeric nanoparticulate drug delivery system: in vitro<br />

characterization of nanoparticles based on sugar-containing conjugates, International Journal of<br />

Pharmaceutics, 245, 67–73, 2002.<br />

9. Kim, S. Y. and Lee, Y. M., Taxol-loaded block copolymer nanospheres composed of methoxy<br />

poly(ethylene glycol) and poly(3-caprolactone) as novel anticancer drug carriers, Biomaterials,<br />

22, 1697–1704, 2001.<br />

10. Kwon, G. S. and Okano, T., Polymeric micelles as new drug carriers, Advanced Drug Delivery<br />

Review, 21, 107–116, 1996.<br />

11. Creutz, S., van Stam, J., De Schryver, F. C., and Jerome, R., Dynamics of poly((dimethylamino)alkyl<br />

methacrylate-block-sodium methacrylate) micelles. Influence of hydrophobicity and molecular<br />

architecture on the exchange rate of copolymer molecules, Macromolecules, 31, 681–689, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 405<br />

12. Zhang, Y., Jin, T., and Zhuo, R., Methotrexate-loaded biodegradable polymeric micelles: Preparation,<br />

physicochemical properties and in vitro drug release, Colloids and Surfaces B. Biointerfaces.<br />

44, 104–109, 2005.<br />

13. Yokoyama, M., Fukushima, S., Uehara, R., Okamoto, K., Kataoka, K., Sakurai, Y., and Okano, T.,<br />

Characterization of physical entrapment and chemical conjugation of adriamycin in polymeric<br />

micelles and their design for in vivo delivery to a solid tumor, Journal of Controlled Release, 50,<br />

79–92, 1998.<br />

14. Yokoyama, M., Opanasopit, P., Okano, T., Kawano, K., and Maitani, Y., Polymer design and incorporation<br />

methods for polymeric micelle carrier system containing water-insoluble anti-cancer agent<br />

camptothecin, Journal of Drug Targeting, 12, 373–384, 2004.<br />

15. You, L.-C., Lu, F.-Z., Li, Z.-C., Zhang, W., and Li, F. -M, Glucose-sensitive aggregates formed <strong>by</strong><br />

poly(ethylene oxide)-block-poly(2-glucosyloxyethyl acrylate) with concanavalin A in dilute aqueous<br />

medium, Macromolecules, 36, 1–4, 2003.<br />

16. Harada, A. and Kataoka, K., Formation of polyion complex micelles in an aqueous milieu from a pair<br />

of oppositrly-charged block copolymers with poly(ethylene glycol) segments, Macromolecules, 28,<br />

5294–5299, 1995.<br />

17. Harada, A. and Kataoka, K., Novel polyion complex micelles entrapping enzyme molecules in the<br />

core: Preparation of narrowly-distributed micelles from lysozyme and poly(ethylene glycol)–poly<br />

(aspartic acid) block copolymers in aqueous medium, Macromolecules, 31, 288–294, 1998.<br />

18. Suh, H., Jeong, B., Rathi, R., and Kim, S. W., Regulation of smooth muscle cell proliferation using<br />

paclitaxel-loaded poly(ethylene oxide)–poly(lactide/glycolide) nanospheres, Journal of Biomedical<br />

Research Materials, 42, 331–338, 1998.<br />

19. Soga, O., van Nostrum, C. F.., Fens, M., Rijcken, C. J. F., Schiffelers, R. M., Storm, G., and Hennink,<br />

W. E., Thermosensitive and biodegradable polymeric micelles for paclitaxel delivery, Journal of<br />

Controlled Release, 103, 341–353, 2005.<br />

20. Burt, H. M., Zhang, X., Toleikis, P., Embree, L., and Hunter, W. L., Development of copolymers of<br />

poly(D,L-lactide) and methoxypolyethylene glycol as micellar carriers of paclitaxel, Colloids and<br />

Surfaces B: Biointerfaces, 16, 161–171, 1999.<br />

21. Zhang, X., Jackson, J. K., and Burt, H. M., Development of amphiphilic diblock copolymers as<br />

micellar carriers of taxol, International Journal of Pharmaceutics, 132, 195–206, 1996.<br />

22. Balasubramanian, D., Srinivas, V., Gaikar, V. G., and Sharma, M. M., Aggregation behavior of<br />

hydrotropic compounds in aqueous solution, Journal of Physical Chemistry, 93, 3865–3870, 1989.<br />

23. Srinivas, V. and Balasubramanian, D., Proline is a protein-compatible hydrotrope, Langmuir,<br />

11, 2830–2833, 1995.<br />

24. Srinivas, V. and Balasubramanian, D., When does the switch from hydrotropy to micellar behavior<br />

occur?, Langmuir, 14, 6658–6661, 1998.<br />

25. Lee, J., Lee, S. C., Acharya, G., Chang, C.-J., and Park, K., Hydrotropic solubilization of paclitaxel:<br />

Analysis of chemical structures for hydrotropic property, Pharmaceutical Research, 20, 1022–1030,<br />

2003.<br />

26. Lee, S. C., Acharya, G., Lee, J., and Park, K., Hydrotropic polymers: Synthesis and characterization of<br />

polymers containing picolylnicotinamide moieties, Macromolecules, 36, 2248–2255, 2003.<br />

27. Huh, K. M., Lee, S. C., Cho, Y. W., Lee, J., Jeong, J. H., and Park, K., Hydrotropic polymer micelle<br />

system for delivery of paclitaxel, Journal of Controlled Release, 101, 59–68, 2005.<br />

28. Gadelle, F., Koros, W. J., and Schechter, R. S., Solubilization of aromatic solutes in block copolymers,<br />

Macromolecules, 28, 4883–4892, 1995.<br />

29. Gan, Z., Jim, T. F., Li, M., Yuer, Z., Wang, S., and Wu, C., Enzymatic biodegradation of poly<br />

(ethylene oxide-b-3-caprolactone) diblock copolymer and its potential biomedical applications,<br />

Macromolecules, 32, 590–594, 1999.<br />

30. Alexandridis, P., Holzwarth, J. F., and Hatton, T. A., Micellization of poly(ethylene oxide)–poly<br />

(propylene oxide)–poly(ethylene oxide) triblock copolymers in aqueous solutions: Thermodynamics<br />

of copolymer association, Macromolecules, 27, 2414–2425, 1994.<br />

31. Kabanov, A. V., Slepnev, V. I., Kuzetsova, L. E., Batrakova, E. V., Alakhov, V. Y., Melik-Nubarov,<br />

N. S., Sveshinikov, P. G., and Kabanov, V. A., Pluronic micelles as tool for low-molecular compound<br />

vector delivery into a cell: Effect of Staphylococcus aureus enterotoxin B on cell loading with micelle<br />

incorporated fluorescent dye, Biochemistry International, 26, 1035–1042, 1992.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


406<br />

Nanotechnology for Cancer Therapy<br />

32. La, S. B., Okano, T., and Kataoka, K., Preparation and characterization of the micelle-forming<br />

polymeric drug indomethacin-incorporated poly(ethylene oxide)–poly(benzyl-2 0 -aspartate) block<br />

copolymer micelles, Journal of Pharmaceutical Sciences, 85, 85–90, 1996.<br />

33. Kalyanasundaram, K. and Thomas, J. K., Environmental effects on vibronic band intensities in pyrene<br />

monomer fluorescence and their application in studies of micellar systems, Journal of the American<br />

Chemical Society, 99, 2039–2044, 1977.<br />

34. Jeong, B., Bae, Y. H., and Kim, S. W., Biodegradable thermosensitive micelles of PEG–PLGA–PEG<br />

triblock copolymers, Colloids and Surfaces B: Biointerfaces, 16, 185–193, 1999.<br />

35. Huh, K. M., Lee, K. Y., Kwon, I. C., Kim, Y.-H., Kim, C., and Jeong, S. Y., Synthesis of triarmed<br />

poly(ethylene oxide)–deoxycholic acid conjugate and its micellar characteristics, Langmuir, 16,<br />

10566–10568, 2000.<br />

36. Kabanov, A. V., Nazarova, I. R., Astafieva, I. V., Batrakova, E. V., Alakhov, V. Y., Yaroslavov, A. A.,<br />

and Kabanov, V. A., Micelle formation and solubilization of fluorescent probes in poly(oxyethyleneb-oxypropylene-b-oxyethylene)<br />

solutions, Macromolecules, 28, 2303–2314, 1995.<br />

37. Lee, K. Y., Jo, W. H., Kwon, I. C., Kim, Y.-H., and Jeong, S. Y., Structural determination and interior<br />

polarity of self-aggregates prepared from deoxycholic acid-modified chitosan in water, Macromolecules,<br />

31, 378–383, 1998.<br />

38. Kim, C., Lee, S. C., Kwon, I. C., Chung, H., and Jeong, S. Y., Complexation of poly(2-ethyl-2oxazoline)-block-poly(3-caprolactone)<br />

micelles with multifunctional carboxylic acids, Macromolecules,<br />

35, 193–200, 2002.<br />

39. Xu, R., Winnik, M. A., Hallett, F. R., Riess, G., and Croucher, M. D., Light-scattering study of the<br />

association behavior of styrene–ethylene oxide block copolymers in aqueous solution, Macromolecules,<br />

24, 87–93, 1991.<br />

40. Ringsdorf, H., Venzmer, J., and Winnik, F. M., Fluorescence studies of hydrophobically modified<br />

poly(N-isopropylacrylamide), Macromolecules, 24, 1678–1686, 1991.<br />

41. Wilhelm, M., Zhao, C.-L., Wang, Y., Xu, R., and Winnik, M. A., Poly(styrene–ethylene oxide) block<br />

copolymer micelle formation in water: a fluorescence probe study, Macromolecules, 24, 1033–1040,<br />

1991.<br />

42. Zhang, L. and Eisenberg, A., Multiple morphologies and characteristics of “crew-cut” micelle-like<br />

aggregates of polystyrene-b-poly(acrylic acid) diblock copolymers in aqueous solutions, Journal of<br />

the American Chemical Society, 118, 3168–3181, 1996.<br />

43. Zhang, L. and Eisenberg, A., Multiple morphologies of “Crew-Cut” aggregates of polystyrene-bpoly(acrylic<br />

acid) block copolymers, Science, 268, 1728–1731, 1995.<br />

44. Zhang, L., Yu, K., and Eisenberg, A., Ion-induced morphological changes in “Crew-Cut” aggregates<br />

of amphiphilic block copolymers, Science, 272, 1777–1779, 1996.<br />

45. Heise, A., Hedrick, J. L., Frank, C. W., and Miller, R. D., Starlike block copolymers with amphiphilic<br />

arms as models for unimolecular micelles, Journal of the American Chemical Society, 121,<br />

8647–8648, 1999.<br />

46. Newkome, G. R., Moorefield, C. N., Baker, G. R., Saunders, M. J., and Grossman, S. H., Unimolecular<br />

micelles, Angewandte Chemie International Edition in English, 30, 1178–1180, 1991.<br />

47. Zhang, G.-D., Harada, A., Nishiyama, N., Jeang, D.-L., Koyama, H., Aida, T., and Kataoka, K.,<br />

Polyion complex micelles entrapping cationic dendrimer porphyrin: Effective photosensitizer for<br />

photodynamic therapy of cancer, Journal of Controlled Release, 93, 141–150, 2003.<br />

48. Jones, M.-C., Ranger, M., and Leroux, J.-C., pH-Sensitive unimolecular polymeric micelles:<br />

Synthesis of a novel drug carrier, Bioconjugate Chemistry, 14, 774–781, 2003.<br />

49. Liu, H., Farrell, S., and Uhrich, K., Drug release characteristics of unimolecular polymeric micelles,<br />

Journal of Controlled Release, 68, 167–174, 2000.<br />

50. Nagasaki, Y., Okada, T., Scholz, C., Iijima, M., Kato, M., and Kataoka, K., The reactive polymeric<br />

micelles based on an aldehyde-ended poly(ethylene glycol)/poly(lactide) block copolymer, Macromolecules,<br />

31, 1473–1479, 1998.<br />

51. Torchilin, V. P., Structure and design of polymeric surfactant-based drug delivery systems, Journal of<br />

Controlled Release, 73, 137–172, 2001.<br />

52. Le Garrec, D., Gori, S., Luo, L., Lessard, D., Smith, D. C., Yessine, M.-A., Ranger, M., and Leroux, J.-<br />

C., Poly(N-vinylpyrrolidone)-block-poly(D,L-lactide) as a new polymeric solubilizer for hydrophobic<br />

anticancer drugs: in vitro and in vivo evaluation, Journal of Controlled Release, 99, 83–101, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Hydrotropic Polymer Micelles for Cancer Therapeutics 407<br />

53. Liggins, R. T. and Burt, H. M., Polyether–polyester diblock copolymers for the preparation of<br />

paclitaxel loaded polymeric micelle formulations, Advanced Drug Delivery Reviews, 54, 191–202,<br />

2002.<br />

54. Liu, S. Q., Tong, Y. W., and Yang, Y. Y., Incorporation and in vitro release of doxorubicin in<br />

thermally sensitive micelles made from poly(N-isopropylacrylamide-co-N,N-dimethylacrylamide)b-poly(D,L-lactide-co-glycolide)<br />

with varying compositions, Biomaterials, 26, 5064–5074, 2005.<br />

55. Lukyanov, A. N. and Torchilin, V. P., Micelles from lipid derivatives of water-soluble polymers as<br />

delivery systems for poorly soluble drugs, Advanced Drug Delivery Reviews, 56, 1273–1289, 2004.<br />

56. U.S. National Institutes of Health. National Cancer Institute, 1986. Developmental Therapeutics<br />

Program. NCI Investigational Drugs: Pharmaceutical Data, pp. 70–72, (http://dtp.nci.nih.gov/docs/<br />

idrugs/chembook.html/).<br />

57. Adams, J. D., Flora, K. P., Goldspiel, B. R., Wilson, J. W., and Arbuck, S. G., Taxol: a history of<br />

pharmaceutical development and current pharmaceutical concerns, Journal of the National Cancer<br />

Institute Monogram, 15, 141–147, 1993.<br />

58. Gref, R., Minamitake, Y., Peracchia, M. T., Trubetskoy, V., Torchilin, V., and Langer, R., Biodegradable<br />

long-circulating polymeric nanospheres, Science, 263, 1600–1603, 1994.<br />

59. Mazzo, D. J., Nguyen-Huu, J.-J., Pagniez, S., and Denis, P., Compatibility of docetaxel and paclitaxel<br />

in intravenous solutions with poyvinyl chloride infusion materials, American Journal of Health-<br />

System Pharmacy, 54, 566–569, 1997.<br />

60. Alkan-Onyuksel, H., Ramakrishnan, S., Chai, H.-B., and Pezzuto, J. M., A mixed micellar formulation<br />

suitable for the parenteral administration of taxol, Pharmaceutical Research, 11, 206–212,<br />

1994.<br />

61. Dordunoo, S. K. and Burt, H. M., Solubility and stability of taxol: Effects of buffer and cyclodextrins,<br />

International Journal of Pharmaceutics, 133, 191–201, 1996.<br />

62. Muller, R. H. and Bohm, B., Nanosuspensions, In Emulsions and Nanosuspensions for the Formulation<br />

of Poorly Soluble Drugs, Muller, Rainer H., Benita, S., and Bohm, B., Eds., Medpharm<br />

Scientific Publishers, Stuttgart, pp. 149–174, 1998.<br />

63. Rubino, J. T. and Yalkowsky, S. H., Cosolvency and cosolvent polarity, Pharmaceutical Research,4,<br />

220–230, 1987.<br />

64. Serajuddin, A. T. M., Solid dispersion of poorly water-soluble drugs: Early promises, subsequent<br />

problems, and recent breakthroughs, Journal of Pharmaceutical Science., 88, 1058–1066, 1999.<br />

65. Sharma, A. and Straubinger, R. M., Novel taxol formulations: Preparation and characterization of<br />

taxol-containing liposomes, Pharmaceutical Research, 11, 889–896, 1994.<br />

66. Tarr, B. D., Sambandan, T. G., and Yalkowsky, S. H., A new parenteral emulsion for the administration<br />

of taxol, Pharmaceutical Research, 4, 162–165, 1987.<br />

67. Nishiyama, N., Bae, Y., Miyata, K., Fukumura, S., and Kataoka, K., Drug Discovery Today:<br />

Technologies, 2, 21–26, 2005.<br />

68. Bae, Y., Nishiyama, N., Fukushima, S., Koyama, H., Yasuhiro, M., and Kataoka, K., Preparation and<br />

biological characterization of polymeric micelle drug carriers with intracellular pH-triggered drug<br />

release property: Tumor permeability, controlled subcellular drug distribution, and enhanced in vivo<br />

antitumor efficacy, Bioconjugate Chemistry, 16, 122–130, 2005.<br />

69. Jaracz, S., Chen, J., Kuznetsova, L. V., and Ojima, I., Recent advances in tumor–targeting anticancer<br />

drug conjugates, Bioorganic & Medicinal Chemistry, 13, 5043–5054, 2005.<br />

70. van Nostrum, C. F., Polymeric micelles to deliver photosensitizers for photodynamic therapy,<br />

Advanced Drug Delivery Reviews, 56, 9–16, 2004.<br />

71. Torchilin, V. P., Polymeric micelles in diagnostic imaging, Colloids and Surfaces B: Biointerfaces,<br />

16, 305–319, 1999.<br />

72. Dolmans, D. and Fukumura, D., Photodynamic therapy for cancer, Nature Reviews Cancer, 3,<br />

380–387, 2003.<br />

73. Yalkowsky, S. H., Solubility and Solubilization in Aqueous Media, American Chemical Society,<br />

Washington, DC, 1999.<br />

74. Gandhi, N. N., Kumar, M. D., and Sathyamurthy, N., Effect of hydrotropes on solubility<br />

and mass-transfer coefficient of butyl acetate, Journal of Chemical & Engineering Data, 43,<br />

695–699, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


408<br />

Nanotechnology for Cancer Therapy<br />

75. Lobenberg, R., Amidon, G. L., and Vierira, M., Solubility as a limiting factor to drug absorption, In<br />

Oral Drug Absorption: Prediction and Assessment, Dressman, J. H. and Lennernas, H., Eds., Marcel<br />

Dekker, New York, pp. 137–153, 2000.<br />

76. Rasool, A. A., Hussain, A. A., and Dittert, L. W., Solubility enhancement of some water-insoluble<br />

drugs in the presence of nicotinamide and related compounds, Journal of Pharmaceutical Science, 80,<br />

387–393, 1991.<br />

77. Coffman, R. E. and Kildsig, D. O., Hydrotropic solubilization-Mechanistic studies, Pharmacal<br />

Research, 13, 1460–1463, 1996.<br />

78. Xia, J., Zhang, X., and Matyjaszewski, K., Atom transfer radical polymerization of 4-vinylpyridine,<br />

Macromolecules, 32, 3531–3533, 1999.<br />

79. Matyjaszewski, K. and Xia, J., Atom transfer radical polymerization, Chemical Reviews, 101,<br />

2921–2990, 2001.<br />

80. Patten, T. E., Xia, J., Abernathy, T., and Matyjaszewski, K., Polymers with very low polydispersities<br />

from atom transfer radical polymerization, Science, 272, 866–868, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


20<br />

CONTENTS<br />

Tumor-Targeted Delivery of<br />

Sparingly-Soluble Anti-Cancer<br />

Drugs with Polymeric Lipid-Core<br />

Immunomicelles<br />

Vladimir P. Torchilin<br />

20.1 Introduction ....................................................................................................................... 409<br />

20.2 Micelles, Micellization, and Drug Delivery ..................................................................... 410<br />

20.3 Tumor Targeting of Polymeric Micelles .......................................................................... 411<br />

References..................................................................................................................................... 417<br />

20.1 INTRODUCTION<br />

Various drug delivery and drug targeting systems-based soluble polymers, nanoparticles made of<br />

natural and synthetic polymers, nanocapsules, lipoproteins, liposomes, micelles, and many other<br />

pharmaceutical carriers are currently widely used to minimize drug degradation and loss upon<br />

administration, prevent harmful or undesirable side-effects, and increase drug bioavailability and<br />

the fraction of the drug accumulated in the required (pathological) zone. 1,2 Many of those drug<br />

carriers are long-circulating 3,4 because prolonged circulation maintains the required therapeutic<br />

level of pharmaceuticals in the blood for extended time intervals, and it allows for slow accumulation<br />

of high-molecular-weight drugs or drug-containing nanocarriers in pathological sites with<br />

affected and leaky vasculature via the enhanced permeability and retention effect (EPR). 5,6 It also<br />

provides better targeting effects for specific ligand-modified drugs and drug carriers. 7<br />

The development of biocompatible, biodegradable, long-circulating, and specifically targeted<br />

drug carriers for sparingly soluble pharmaceuticals that possess high loading capacities is attracting<br />

a lot of attention because the therapeutic application of hydrophobic, poorly water-soluble agents<br />

might be associated with serious problems as low water-solubility results in poor absorption and<br />

low bioavailability upon oral administration. 8 The aggregation of poorly soluble drugs upon intravenous<br />

administration might lead to an embolism 9 and increased local toxicity at the sites of<br />

aggregate deposition. 10 However, the hydrophobicity (low water solubility) allows a drug molecule<br />

to better penetrate a cell membrane when the molecular target for the drug is intracellularly<br />

located, 11,12 and this is especially important for certain anti-cancer agents, many of which are<br />

bulky polycyclic compounds. 13 To overcome some drugs’ poor solubility and to increase their<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

409


410<br />

bioavailability, certain clinically acceptable organic solvents are used in pharmaceutical formulations<br />

such as ethanol or polyethoxylated castor oil (Cremophor EL) 10 as well as drug carriers such<br />

as liposomes 14 and cyclodextrins. 15 In addition, a lot of interest is currently expressed toward using<br />

various micelle-forming substances in formulations of insoluble drugs. 10<br />

20.2 MICELLES, MICELLIZATION, AND DRUG DELIVERY<br />

Nanotechnology for Cancer Therapy<br />

Micelles represent dispersed systems, consisting of particulate matter or dispersed phase that are<br />

distributed within a continuous phase or dispersion medium and, under certain conditions (concentration<br />

and temperature), are spontaneously formed <strong>by</strong> amphiphilic or surface-active agents<br />

(surfactants) that are molecules that consist of two clearly distinct moieties differing in their affinity<br />

toward a given solvent. 16 They typically have particle sizes within a 5 to 50–100 nm range. The<br />

concentration of a monomeric amphiphile where micelles (aggregates, including several dozens of<br />

amphiphilic molecules and usually spherical in shape) appear is called the critical micelle (or<br />

micellization) concentration (CMC). The lower the CMC value of a given amphiphilic polymer,<br />

the more stable micelles are even at low net concentration of an amphiphile in the medium. This is<br />

important from the practical point of view because upon dilution with the large volume of the blood,<br />

only micelles with low CMC value still exist, whereas micelles with high CMC value may dissociate<br />

into uinimers, and their content may precipitate in the blood. In aqueous media,<br />

hydrophobic fragments of amphiphilic molecules form the core of a micelle that can solubilize<br />

poorly soluble pharmaceuticals, 17 whereas polar molecules will be adsorbed on the micelle<br />

surface, and substances with intermediate polarity/solubility will be distributed in various<br />

intermediate positions.<br />

Micellization of biologically active substances is a wide-spread phenomenon that is important<br />

for pharmacological action. For example, the increase in the bioavailability of a lipophilic drug<br />

upon oral administration is attributed to drug solubilization in the gut <strong>by</strong> naturally occurring biliary<br />

lipid/fatty acid-containing mixed micelles produced <strong>by</strong> organism upon the digestion of the dietary<br />

fat. Surfactant micelles are also widely used as adjuvants and drug carrier systems, and in many<br />

areas of pharmaceutical technology and controlled drug delivery research, micelles as drug carriers<br />

provide a set of clear advantages. 18–22 The solubilization of sparingly soluble drugs <strong>by</strong> micelleforming<br />

surfactants results in an increased water solubility of sparingly soluble drug and its<br />

improved bioavailability, extended blood half-life upon intravenous administration, reduction of<br />

toxicity and other adverse effects, enhanced permeability across the physiological barriers, and<br />

substantial changes in drug biodistribution. In addition, being in a micellar container, the drug is<br />

well-protected from possible inactivation under the effect of biological surroundings and does not<br />

provoke toxic side effects on non-target tissues.<br />

Polymeric micelles, especially popular in drug delivery research, are formed from block-copolymers<br />

consisting of hydrophilic and hydrophobic monomer units with the length of a hydrophilic<br />

block exceeding, to some extent, that of a hydrophobic one. These micelles contain the core of the<br />

hydrophobic blocks stabilized <strong>by</strong> the corona of hydrophilic polymeric chains. If the length of a<br />

hydrophilic block is too high, copolymers exist in water as unimers (individual molecules), whereas<br />

molecules with very long hydrophobic block forms structure with non-micellar morphology. 23 The<br />

core compartment of the pharmaceutical polymeric micelle should demonstrate high loading<br />

capacity, controlled release profile for the incorporated drug, and good compatibility between<br />

the core-forming block and incorporated drug. The micelle corona should provide an effective<br />

steric protection for the micelle. It should also determine for the micelle hydrophilicity, charge,<br />

the length and surface density of hydrophilic blocks, and the presence of reactive groups suitable<br />

for further micelle derivatization such as an attachment of targeting moieties. The listed properties<br />

control key biological characteristics of a micellar carrier including pharmacokinetics,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer Drugs 411<br />

biodistribution, biocompatibility, longevity, surface adsorption of biomacromolecules, adhesion to<br />

biosurfaces, and targetability. 1,24–30<br />

PEG still remains the hydrophilic block of choice, although some other hydrophilic polymers<br />

may be used to make corona blocks, 31,32 whereas a variety of polymers was used to build hydrophobic<br />

core-forming blocks including propylene oxide, 26,33 L-lysine, 34,35 aspartic acid, 36,37<br />

b-benzoyl-L-aspartate, 38,39 caprolactone, 40,41 D,L-lactic acid, 27,42 and many others. Phospholipid<br />

residues can also be successfully used as hydrophobic core-forming groups. 43 The use of lipid<br />

moieties as hydrophobic blocks capping hydrophilic polymer (such as PEG) chains can provide<br />

additional advantages for particle stability when compared with conventional amphiphilic polymer<br />

micelles because of the existence of two fatty acid acyls that may considerably contribute to an<br />

increase in the hydrophobic interactions in the micelle’s core. Although diacyllipid–PEG conjugates<br />

have been introduced into the area of controlled drug delivery as polymeric surface modifiers<br />

for liposomes, 44 the diacyllipid–PEG molecule itself represents a typical amphiphilic copolymer<br />

with a bulky hydrophilic (PEG) portion and a very short, but extremely hydrophobic, diacyllipid<br />

part. Similar to other PEG-containing amphiphilic block-copolymers, diacyllipid–PEG conjugates<br />

were found to form micelles in an aqueous environment. 45 A series of PEG–phosphatidylethanolamine<br />

(PEG–PE) conjugates was synthesized using egg PE and N-hydroxysuccinimide esters of<br />

methoxy–PEG succinates with different MW of PEG fragments. 44 HPLC-based gel permeation<br />

chromatography showed that these polymers formed micelles of different sizes in water. The<br />

stability of the polymeric micelles was confirmed <strong>by</strong> the fact that no dissociation into individual<br />

polymeric chains was found following the chromatography of the serially diluted samples of<br />

PEG–PE up to a polymer concentration of ca. 1 ml/ml corresponds to a micromolar CMC value<br />

that is at least 100-fold lower than those of conventional detergents. 46 Such low CMC values<br />

indicate that micelles prepared from these polymers will maintain their integrity even upon<br />

strong dilution (for example, in the blood during a therapeutic application). High stability of<br />

polymeric micelles allows also for good retention of encapsulated drugs in the solubilized form<br />

upon parenteral administration. PEG–PE micelles also can efficiently incorporate a broad variety of<br />

sparingly soluble and amphiphilic substances including paclitaxel, tamoxifen, camptothecin,<br />

porphyrine, vitamin K3, and others. 47–49<br />

Among other factors influencing the efficacy of drug loading into the micelle are the sizes of<br />

both core-forming and corona-forming blocks. 50 In the first case, the larger the hydrophobic block,<br />

the bigger core size and the ability of micelle to entrap hydrophobic drugs. In the second case, the<br />

increase in the length of the hydrophilic block results in the increase of the CMC value, i.e., at a<br />

given concentration of the amphiphilic polymer in solution the smaller fraction of this polymer will<br />

be present in the micellar form and the quantity of the micelle-associated drug drops. Incorporation<br />

into polymeric micelles is beneficial for various sparingly soluble drugs. Therefore, doxorubicin<br />

incorporated into Pluronic w micelles demonstrated superior properties as compared to free drugs in<br />

the experimental treatment of murine tumors (leukemia P388, myeloma, and Lewis lung carcinoma)<br />

and human tumors (breast carcinoma MCF-7) in mice. 51 In addition, the reduction of the side<br />

effects of the drug was observed in many cases. The toxicity decrease (smaller vascular damage and<br />

liver focal necrosis) was found for the polymeric micelle-incorporated anti-tumor drug KRN5500. 52<br />

20.3 TUMOR TARGETING OF POLYMERIC MICELLES<br />

Making micelles capable of specific accumulation in desired areas in the body (tumors) can further<br />

increase the efficiency of micelle-encapsulated pharmaceuticals. The simplest way to achieve this is<br />

to rely on previously mentioned preferential accumulation of drug-loaded micelles in areas with<br />

compromised vasculature (tumors and infarcts) via the EPR effect that is based on the spontaneous<br />

penetration of long-circulating macromolecules, molecular aggregates, and particulate drug carriers<br />

into the interstitium through the leaky blood vessels in certain pathological sites in the body.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


412<br />

It has been shown that the EPR effect is typical for solid tumors and infarcts. 5,6 Micelles prepared<br />

from PEG–PE conjugates with different length of the PEG block demonstrated much higher<br />

accumulation in tumors compared to non-target tissue (muscle) in experimental Lewis lung carcinoma<br />

(<strong>LLC</strong>: tumor with a relatively low vascular permeability) 53,54 in mice. The peak uptake is<br />

achieved around five hours post-injection, and the maximal total tumor uptake of the injected dose<br />

within the observation period (AUC) was found for micelles made PEG–PE with the largest size of<br />

the PEG block (5,000 Da). This was explained <strong>by</strong> the fact that these micelles had little extravasation<br />

into the normal tissue compared to micelles prepared from the shorter PEG–PE conjugates.<br />

Micelles prepared from PEG–PE conjugates with shorter PEG, however, might be more efficient<br />

carriers of poorly soluble drugs because they have a greater hydrophobic-to-hydrophilic phase ratio<br />

and can be loaded with drugs more efficiently on a weight-to-weight basis. Similar results were also<br />

obtained in another murine tumor model, EL4 T cell lymphoma (EL4); 55 see Figure 20.1.<br />

It is important that tumor diffusion and accumulation parameters of nanosized pharmaceutical<br />

carriers strongly depend on the cutoff size of tumor blood vessel wall, and the cutoff size varies for<br />

different tumors. 56–58 Because tumor vasculature permeability depends on the particular type of the<br />

tumor, 57 the use of micelles as drug carriers could be specifically justified for tumors whose<br />

vasculature has the low cutoff size (below 200 nm). The use of PEG–PE micelles for the effective<br />

delivery of a model protein drug (soybean trypsin inhibitor or STI, MW 21.5 kDa) to subcutaneously<br />

established <strong>LLC</strong> in mice was reported. 54<br />

An interesting targeting mechanism is based on the fact that many pathological processes in<br />

various tissues and organs are accompanied with local temperature increase and/or acidosis. 59,60<br />

Micelles made capable of disintegration under the increased temperature or decreased pH values in<br />

pathological sites (such as tumors) can combine the EPR effect-mediated accumulation with<br />

stimuli-responsiveness. It was shown that micelles made of thermo- or pH-sensitive components<br />

such as poly(N-isopropylacrylamide) and its copolymers with poly(D,L-lactide) and other blocks<br />

acquire the ability to disintegrate in target areas, releasing the micelle-incorporated drug. 61–65<br />

pH-responsive polymeric micelles loaded with phtalocyanine seem to be promising carriers for<br />

the photodynamic cancer therapy, 66 whereas doxorubicin-loaded polymeric micelles containing<br />

acid-cleavable linkages provided an enhanced intracellular drug delivery into tumor cells and,<br />

therefore, higher efficiency. 67<br />

As with many other delivery systems, the drug delivery potential of polymeric micelles may be<br />

further enhanced <strong>by</strong> attaching targeting ligands to their surface. The attachment of various specific<br />

ligand to the water-exposed termini of hydrophilic blocks has been used to improve the targeting<br />

of micelles and micelle-incorporated drugs and DNA. 26,68,69 Among those ligands, various sugar<br />

moieties can be named as well as 70 transferrin, 71 and folate residues 72 because many target cells,<br />

AUC dose/g organ<br />

75<br />

50<br />

25<br />

Tumor<br />

Muscle<br />

Tumor<br />

Muscle<br />

0<br />

0<br />

(a) PEG750-PE PEG2000-PE (b) PEG750-PE PEG2000-PE FIGURE 20.1 Accumulation of PEG–PE micelles in tumors in mice compared to muscle accumulation.<br />

(a) <strong>LLC</strong> tumor; (b) EL4 tumor.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

AUC dose/g organ<br />

75<br />

50<br />

25<br />

Nanotechnology for Cancer Therapy


Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer Drugs 413<br />

especially cancer cells, over express appropriate receptors (such as transferrin and folate receptors)<br />

on their surface. Transferrin-modified micelles based on PEG and polyethyleneimine with sizes<br />

between 70 and 100 are expected to target tumors with over expressed transferrin receptors. 71<br />

Similar targeting approaches were successfully tested with folate-modified micelles. 73 Poly(Lhistidine)/PEG<br />

and poly(L-lactic acid)/PEG block copolymer micelles carrying folate residue on<br />

their surfaces were shown to be efficient for the delivery of adriamycin to tumor cells in vitro,<br />

demonstrating potential for solid tumor treatment and combined targetability and pH-sensitivity. 74<br />

Among all specific ligands, monoclonal antibodies, primarily of the IgG class, and their Fab<br />

fragments provide the broadest opportunities in terms of diversity of targets and specificity of<br />

interaction. Several successful attempts to covalently attach an antibody to a surfactant or polymeric<br />

micelles (i.e., to prepare immunomicelles) have been already described. 22,26,71,75 By adapting<br />

the coupling technique developed for attaching specific ligands to liposomes utilizing PEG–PE with<br />

the free PEG terminus activated with p-nitrophenylcarbonyl (pNP) group, 76 PEG–PE-based immunomicelles<br />

have been prepared modified with monoclonal antibodies (see the principal reaction<br />

scheme in Figure 20.2). The micelle-attached protein was quantified using fluorescent labels or <strong>by</strong><br />

SDS-PAGE, 75,77 and it was calculated that 10–20 antibody molecules could be attached to a single<br />

micelle. Antibodies attached to the micelle corona preserve their specific binding ability, and<br />

immunomicelles specifically recognize their target substrates as was confirmed <strong>by</strong> ELISA experiments<br />

while the micelles maintain essentially the same size as non-modified ones. 75<br />

One has to keep in mind, however, that whole antibody attachment to the micelles (as well as to<br />

any other drug carriers) might provoke faster clearance from the circulation as a result of uptake <strong>by</strong><br />

Fc receptor-bearing Kupffer cells. To test if the antibody attachment affects the blood clearance of<br />

PEG–PE micelles, the comparison was made of the clearance characteristics of plain and antibodymodified<br />

micelles in mice; it was found that PEG–PE micelle modification with an antibody had a<br />

very small effect on their blood clearance. 75<br />

To specifically enhance the tumor accumulation of PEG–PE-based micelles, they have been<br />

modified with tumor-specific monoclonal antibodies. 75,77 Non-pathogenic monoclonal antinuclear<br />

autoantibodies with the nucleosome-restricted specificity (monoclonal antibody 2C5, mAb 2C5<br />

being among them) that recognize the surfaces of numerous tumor, but not normal, cells via tumor<br />

cell surface-bound nucleosomes 78,79 were used in these experiments. Because these antibodies bind<br />

a broad variety of cancer cells, they may serve as specific ligands for the delivery of drugs and drug<br />

carriers into tumors. The data shown in Figure 20.3a indicate that fluorescence (rhodamine)-labeled<br />

2C5-immunomicelles effectively bind to the surface of several unrelated human and murine tumor<br />

cells lines: human BT20 (breast adenocarcinoma), murine <strong>LLC</strong> (Lewis lung carcinoma), and EL4<br />

(T lymphoma) cells. The incubation of antibody-free micelles with the same cells results in<br />

virtually no micelle-to-cell association. Drug(paclitaxel)-loaded 2C5-immunomicelles also effectively<br />

bound various tumors cells (Figure 20.3b). Such specific recognition of cancer cells <strong>by</strong> drugloaded<br />

mAb 2C5-immunomicelles results in dramatically improved cancer cell killing <strong>by</strong> such<br />

micelles. Figure 20.3c presents the results of in vitro experiments with human breast cancer MCF-7<br />

cells that showed a clearly superior efficiency of paclitaxel-loaded 2C5-immunomicelles compared<br />

to paclitaxel-loaded plain or modified with a non-specific IgG micelles or free drug.<br />

In vivo experiments with <strong>LLC</strong> tumors-bearing mice revealed a dramatically enhanced tumor<br />

uptake of paclitaxel-loaded radiolabeled 2C5-immunomicelles compared to non-targeted micelles<br />

(Table 20.1). For tumor accumulation experiments in vivo, <strong>LLC</strong> tumors have been grown in mice<br />

subcutaneously injected with 20,000 <strong>LLC</strong> cells in 50 mL of 10 mM HBS into the left rear flank.<br />

When tumor diameters reached 3–7 mm (8–12 days post inoculation), the mice were injected with<br />

111 In-labeled plain or mAb 2C5-bearing paclitaxel-loaded PEG–PE micelles via the tail vein. At<br />

30 min or 2 h post injection, mice were sacrificed; tumors and muscle samples were collected and<br />

analyzed for the presence of 111 In radioactivity to estimate micelle accumulation in the tumor. To<br />

estimate the actual tumor accumulation of the drug, samples of tumor tissues from mice receiving<br />

ca. 100 mg paclitaxel per animal in plain PEG–PE micelles, in 2C5-immunomicelles, and the same<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


O 2 N<br />

O<br />

C O<br />

O 2 N<br />

O<br />

C O<br />

O<br />

O<br />

C<br />

NO 2<br />

O<br />

O C<br />

Ligand<br />

NO2 HO<br />

N2H FIGURE 20.2 Schematic pattern of antibody (or other amino-containing ligand) attachment to activated PEG–PE micelles via the pNP group.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

OH<br />

414<br />

Nanotechnology for Cancer Therapy


Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer Drugs 415<br />

BT-20<br />

<strong>LLC</strong><br />

EL4<br />

BT-20<br />

<strong>LLC</strong><br />

EL4<br />

(a) (b) (c)<br />

1 2 3<br />

quantity of paclitaxel in Cremophor EL/ethanol/saline mixture (free paclitaxel) were analyzed<br />

30 min and 2 h post injection <strong>by</strong> HPLC. The results presented in the data in the Table 20.1<br />

clearly demonstrate that paclitaxel-loaded mAb 2C5-immunomicelles accumulate in tumors significantly<br />

better than drug-loaded plain micelles or free drugs and bring into tumors more drug<br />

(paclitaxel) than any other preparation. An enhanced accumulation of 2C5-targeted micelles over<br />

plain micelles in the tumor (up to 30%) was observed both at 30 min and at 2 h post injection<br />

indicating the specific recognition and tumor binding of 2C5-targeted immunomicelles. These data<br />

suggest the possibility that drug-loaded immunomicelles may also be better internalized <strong>by</strong> tumor<br />

cells similar to antibody-targeted liposome, 80 delivering more drug inside tumor cells that might be<br />

achieved in case of simple EPR effect-mediated tumor accumulation. By analyzing the absolute<br />

quantity of tumor-accumulated paclitaxel (HPLC 81 ) delivered <strong>by</strong> different drug formulations, it was<br />

shown that 2C5-immunomicelle was bringing into tumors substantially higher quantities of paclitaxel<br />

than in the case of paclitaxel-loaded non-targeted micelles or free drug formulation<br />

(Taxol w )(see Table 20.1). The difference in tumor accumulation of paclitaxel in immunomicelles<br />

compared to other drug preparations was larger at 2 h post injection than at 30 min post injection,<br />

explained <strong>by</strong> the accumulation of different preparations in different tumor compartments. Nontargeted<br />

paclitaxel formulations accumulate in the interstitial space of the tumor via the EPR effect<br />

Viable cells (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

free paclitaxel<br />

plain micelles<br />

2C5-immunomicelles<br />

FIGURE 20.3 Specific properties of PEG–PE-based 2C5-immunomicelles in vitro. (a) Enhanced binding of<br />

rhodamine-labeled 2C5-immunomicelles with various cancer cells. (b) Enhanced binding of rhodaminelabeled<br />

drug(paclitaxel)-loaded 2C5-immunomicelles with various cancer cells. (c) Increased cytotoxicity<br />

of paclitaxel-loaded 2C5-immunomicelles toward cancer cells (with MCF7 cells as an example).<br />

TABLE 20.1<br />

Paclitaxel-Loaded PEG–PE-Based Micelles In Vivo<br />

Paclitaxel in Pain Micelles Paclitaxel in 2C5-Micelles Free Paclitaxel<br />

Tumor accumulation of<br />

micelles (% dose/g of tumor)<br />

4.15G0.36 6G0.10 n/a<br />

Tumor accumulation of<br />

paclitaxel (ng/g of tumor)<br />

200G30 850G40 130G25<br />

Inhibition of tumor growth (%) 35G5 74G34 23G4<br />

Accumulation measured 2 h post injection. Tumor growth inhibition measured over 10 days. Immunomicelles show the<br />

highest accumulation of all preparations, bring the maximum quantity of the drug into tumor, and allow for maximum<br />

inhibition of tumor growth.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


416<br />

and eventually drug could be cleared from the site after gradual micelle degradation. Contrary to<br />

that, paclitaxel-loaded 2C5-immunomicelles are internalized <strong>by</strong> cancer cells and keep the drug<br />

inside the tumor in a way similar to what was observed with drug-loaded anti-her2 immunoliposomes.<br />

80 The internalization <strong>by</strong> tumor cells may be highly therapeutically significant for many antitumor<br />

agents. For example, a much higher tumor regression was observed with a carrier capable of<br />

intracellular drug delivery for an equal doxorubicin dose delivered to the tumor. 80 It results in a<br />

higher therapeutic efficiency of paclitaxel-loaded mAb 2C5-micelles. The average weight of<br />

excised tumors in the group treated with paclitaxel in mAb 2C5-immunomicelles was approx.<br />

0.7 g compared to 1.6 g and 1.4 g in groups treated with the same quantity of paclitaxel as free<br />

drug or in plain PEG–PE micelles, respectively, (P!.05 in both cases). The weight of untreated<br />

tumors was around 2.0 g (Table 20.1).<br />

Recently, a poorly soluble photodynamic therapy (PDT) agent, meso-tetratphenylporphine<br />

(TPP), was effectively solubilized using non-targeted and tumor-targeted polymeric micelles<br />

prepared with PEG–PE. 82 Encapsulation of TPP into PEG–PE-based micelles and 2C5-immunomicelles<br />

(bearing an anti-cancer monoclonal 2C5 antibody) resulted in significantly improved anticancer<br />

effects of the drug at PDT conditions against murine (<strong>LLC</strong>, B16) and human (MCF-7, BT20)<br />

cancer cells in vitro. For this purpose, the cells were incubated for 6 h or 18 h with the TPP or TPPloaded<br />

PEG–PE micelles/immunomicelles and then light-irradiated for 30 min. The phototoxic<br />

effect depended on the TPP concentration and the light dose (the duration of light-irradiation)<br />

(see some data in Figure 20.4). An increased level of apoptosis was shown in the PDT-treated<br />

cultures. The attachment of the anti-cancer 2C5 antibodies to TPP-loaded micelles provided the<br />

maximum level of cell killing at a given time. Therefore, drug-loaded immunomicelles may also<br />

represent a useful formulation of the photosensitizer for practical PDT.<br />

Certain polymeric micelles possess a very high stability and ability to carry a variety of poorly<br />

soluble pharmaceuticals, and they can be used as targeted drug delivery systems into tumors. The<br />

tumor targeting can be achieved via the EPR effect <strong>by</strong> making micelles of stimuli-responsive<br />

amphiphilic block-copolymers or <strong>by</strong> attaching specific targeting ligand molecules (such as folate<br />

of transferrin) to the micelle surface. Tumor-specific immunomicelles can be prepared <strong>by</strong> coupling<br />

monoclonal antibody (such as antinucleosomal monoclonal antibody) molecules to pNP groups on<br />

the water-exposed terimini of the micelle corona-forming blocks. The micelle-coupled antibodies<br />

preserve their specific activity, and immunomicelles prepared with the use of the cancer-specific<br />

mAb 2C5 specifically bind to different cancer cells in vitro, demonstrating increased accumulation<br />

% survival<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0<br />

10 20 30<br />

TPP concentration (μg/ml)<br />

Nanotechnology for Cancer Therapy<br />

40 50<br />

FIGURE 20.4 Improved killing of cancer cells with the drug loaded into tumor-specific 2C5-immunomicelles<br />

under the PDT conditions. Phototoxicity of free TPP and TPP encapsulated in PEG–PE micelles and 2C5-<br />

PEG–PE-immunomicelles after light irradiation towards BT20 (human breast adenocarcinoma) cells; all<br />

preparations were pre-incubated with cells for 18 h): Cfree TPP; $TPP in PEG–PE micelles; %TPP in<br />

2C5-PEG–PE-immunomicelles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer Drugs 417<br />

in experimental tumors in vivo. Being loaded with poorly soluble anti-cancer drugs (such as<br />

paclitaxel, camptothecin, 83 or photodynamic therapy agents), 82 mAb 2C5-immunomicelles demonstrate<br />

significantly increased cytotoxicity toward tumor cells in vitro and in vivo.<br />

REFERENCES<br />

1. Müller, R. H., Colloidal Carriers for Controlled Drug Delivery and Targeting, Wissenschaftliche<br />

Verlagsgesellschaft/CRC Press, Stuttgart, Germany/Boca Raton, 1991.<br />

2. Cohen, 2. S. and Bernstein, H., Eds., Microparticulate Systems for the Delivery of Proteins and<br />

Vaccines, Marcel Dekker, Inc., New York, 1996.<br />

3. Lasic, D. D. and Martin, F., Eds., Stealth Liposomes, CRC Press, Boca Raton, 1995.<br />

4. Torchilin, V. P. and Trubetskoy, V. S., Which polymers can make nanoparticulate drug carriers longcirculating?,<br />

Adv. Drug Deliv. Rev., 16, 141, 1995.<br />

5. Palmer, T. N. et al., The mechanism of liposome accumulation in infarction, Biochim. Biophys. Acta,<br />

797, 363, 1984.<br />

6. Maeda, H. et al., Tumor vascular permeability and the EPR effect in macromolecular therapeutics: A<br />

review, J. Control. Rel., 65, 271, 2000.<br />

7. Torchilin, V. P., Polymer-coated long-circulating microparticular pharmaceuticals, J. Microencapsul.,<br />

15, 1, 1998.<br />

8. Lipinski, C. A. et al., Experimental and computational approaches to estimate solubility and permeability<br />

in drug discovery and development settings, Adv. Drug Deliv. Rev., 46, 3, 2001.<br />

9. Fernandez, A. M. et al., N-Succinyl-(beta-alanyl-L-leucyl-L-alanyl-L-leucyl)doxorubicin: An extracellularly<br />

tumor-activated prodrug devoid of intravenous acute toxicity, J. Med. Chem., 44, 3750,<br />

2001.<br />

10. Yalkowsky, S. H., Ed,. Techniques of Solubilization of Drugs, Marcel Dekker, New York, 1981.<br />

11. Yokogawa, K. et al., Relationships in the structure-tissue distribution of basic drugs in the rabbit,<br />

Pharm. Res., 7, 691, 1990.<br />

12. Hageluken, A. et al., Lipophilic beta-adrenoceptor antagonists and local anesthetics are effective<br />

direct activators of G-proteins, Biochem. Pharmacol., 47, 1789, 1994.<br />

13. Shabner, B. A. and Collings, J. M., Eds., Cancer Chemotherapy: Principles and Practice, J.B.<br />

Lippincott Co., Philadelphia, 1990.<br />

14. Lasic, D. D. and Papahadjopoulos, D., Medical Applications of Liposomes, Elsevier, New York, 1998.<br />

15. Thompson, D. and Chaubal, M. V., Cyclodextrins (CDS)—excipients <strong>by</strong> definition, drug delivery<br />

systems <strong>by</strong> function (part I: Injectable applications), Drug Deliv. Technol., 2, 34, 2000.<br />

16. Mittal, K. L. and Lindman, B., Eds., Surfactants in Solution, Vol. 1–3, Plenum Press, New York, 1991.<br />

17. Lasic, D. D., Mixed micelles in drug delivery, Nature, 355, 279, 1992.<br />

18. Kwon, G. S. and Kataoka, K., Block copolymer micelles as long-circulating drug vehicles, Adv. Drug<br />

Deliv. Rev., 16, 295, 1995.<br />

19. Kwon, G. S., Diblock copolymer nanoparticles for drug delivery, Crit. Rev. Ther. Drug Carrier Syst.,<br />

15, 481, 1998.<br />

20. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., Pluronic block copolymers as novel polymer<br />

therapeutics for drug and gene delivery, J. Control. Rel., 82, 189, 2002.<br />

21. Jones, M. and Leroux, J., Polymeric micelles-a new generation of colloidal drug carriers, Eur.<br />

J. Pharm. Biopharm., 48, 101, 1999.<br />

22. Torchilin, V. P., Structure and design of polymeric surfactant-based drug delivery systems, J. Control.<br />

Rel., 73, 137, 2001.<br />

23. Zhang, L. and Eisenberg, A., Multiple morphologies of “crew-cut” aggregates of polystyrene-bpoly(acrylic<br />

acid) block copolymers, Science, 268, 1728, 1995.<br />

24. Gref, R. et al., Biodegradable long-circulating polymeric nanospheres, Science, 263, 1600, 1994.<br />

25. Gref, R. et al., The controlled intravenous delivery of drugs using PEG-coated sterically stabilized<br />

nanospheres, Adv. Drug Deliv. Rev., 16, 215, 1995.<br />

26. Kabanov, A. V. et al., The neuroleptic activity of haloperidol increases after its solubilization in<br />

surfactant micelles, FEBS Lett., 258, 343, 1989.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


418<br />

Nanotechnology for Cancer Therapy<br />

27. Hagan, S. A. et al., Polylactide-poly(ethelene glycol) copolymers as drug delivery systems 1. Caracterization<br />

of water dispersible micelle-forming systems, Langmuir, 12, 2153, 1996.<br />

28. Inoue, T. et al., An AB block copolymers of oligo(methyl methacrylate) and poly(acrylic acid) for<br />

micellar delivery of hydrophobic drugs, J. Control. Rel., 51, 221, 1998.<br />

29. Hunter, R. J., Foundations of Colloid Science, Vol. 1, Oxford University Press, New York, 1991.<br />

30. Kuntz, R. M. and Saltzman, W. M., Polymeric controlled delivery for immunization, Trends Biotech.,<br />

15, 364, 1997.<br />

31. Torchilin, V. P. et al., Amphiphilic vinyl polymers effectively prolong liposome circulation time<br />

in vivo, Biochim. Biophys. Acta, 1195, 181, 1994.<br />

32. Torchilin, V. P. et al., New sunthetic amphiphilic polymers for steric protection of liposomes in vivo,<br />

J. Pharm. Sci., 84, 1049, 1995.<br />

33. Miller, D. W. et al., Interactions of pluronic block copolymers with brain microvessel endothelial<br />

cells: Evidence of two potential pathways for drug absorption, Bioconjug. Chem., 8, 649, 1997.<br />

34. Katayose, S. and Kataoka, K., Remarkable increase in nuclease resistance of plasmid DNA through<br />

supramolecular assembly with poly(ethylene glycol)–poly(L-lysine) block copolymer, J. Pharm. Sci.,<br />

87, 160, 1998.<br />

35. Trubetskoy, V. S. et al., Block copolymer of polyethylene glycol and polylysine as a carrier of organic<br />

iodine: Design of a long circulating particulate contrast medium for x-ray computed tomography,<br />

J. Drug Target., 4, 381, 1997.<br />

36. Yokoyama, M. et al., Characterization and anticancer activity of the micelle-forming polymeric<br />

anticancer drug adriamycin-conjugated poly(ethylene glycol)–poly(aspartic acid) block copolymer,<br />

Cancer Res., 50, 1693, 1990.<br />

37. Harada, A. and Kataoka, K., Novel polyion complex micelles entrapping enzyme molecules in the<br />

core. Preparation of narrowly-distributed micelles from lysozyme and poly(ethylene glycol)–poly<br />

(aspartic acid) block copolymer in aqueous medium, Macromolecules, 31, 288, 1998.<br />

38. La, S. B., Okano, T., and Kataoka, K., Preparation and characterization of the micelle-forming<br />

polymeric drug indomethacin-incorporated poly(ethylene oxide)–poly(-benzyl L-aspartat) block<br />

copolimer micelles, J. Pharm. Sci., 85, 85, 1996.<br />

39. Kwon, G. S. et al., Block copolymer micelles for drug delivery: Loading and release of doxorubicin,<br />

J. Control. Rel., 48, 195, 1997.<br />

40. Kim, S. Y. et al., Metoxy poly(ethylene glucol) and 3-caprolactone amphiphilic block copolymeric<br />

micelle containing indomethacin, II. Micelle formation and drug release behaviors, J. Control. Rel.,<br />

51, 13, 1998.<br />

41. Allen, C. et al., Polycaprolactone-b-poly(ethylene oxide) block copolymer micelles as a novel drug<br />

delivery vehicle for neurotrophic agents FK506 and L-685,818, Bioconjug. Chem., 9, 564, 1998.<br />

42. Ramaswamy, M. et al., Human plasma distribution of free paclitaxel and paclitaxel associated with<br />

diblock copolymers, J. Pharm. Sci., 86, 460, 1997.<br />

43. Trubetskoy, V. S. and Torchilin, T. P., Use of polyoxyethylene–lipid conjugates as long-circulating<br />

carriers for delivery of therapeutic and diagnostic agents, Adv. Drug Deliv. Rev., 16, 311, 1995.<br />

44. Klibanov, A. L. et al., Amphipathic polyethyleneglycols effectively prolong the circulation time of<br />

liposomes, FEBS Lett., 268, 235, 1990.<br />

45. Lasic, D. D. et al., Valentincic, phase behavior of “stealth–lipid” decithin mixtures, Period. Biol., 93,<br />

287, 1991.<br />

46. Ray, R. et al., Handbook of Pharmaceutical Excipients, APhA Publications, Washington, 2003.<br />

47. Trubetskoy, V. S. and Torchilin, V. P., Polyethylene glycol based micelles as carriers of therapeutic<br />

and diagnostic agents, S.T.P. Pharma Sci., 6, 79, 1996.<br />

48. Gao, Z. et al., Diacyl-polymer micelles as nanocarriers for poorly soluble anticancer drugs, Nano<br />

Lett., 2, 979, 2002.<br />

49. Wang, J. et al., Preparation and in vitro synergistic anticancer effect of Vitamin K# and 1,8-diazabicyclo[5,4,0]undec-7-ene<br />

in poly(ethylene glycol)-diacyllipid micelles, Int. J. Pharm., 272, 129, 2004.<br />

50. Allen, C., Maysinger, D., and Eisenberg, A., Nano-engineering block copolymer aggregates for drug<br />

delivery, Coll. Surf. B: Biointerf., 16, 1, 1999.<br />

51. Alakhov, V. Yu. and Kabanov, A. V., Block copolymeric biotransport carriers as versatile vehicles for<br />

drug delivery, Expert. Opin. Invest. Drugs, 7, 1453, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Tumor-Targeted Delivery of Sparingly-Soluble Anti-Cancer Drugs 419<br />

52. Matsumura, Y. et al., Reduction of the side effects of an antitumor agent, KRN5500, <strong>by</strong> incorporation<br />

of the drug into polymeric micelles, Jpn. J. Cancer Res., 90, 122, 1999.<br />

53. Wang, J., Mongayt, D., and Torchilin, V. P., Polymeric micelles for delivery of poorly soluble drugs:<br />

Preparation and anticancer activity in vitro of paclitaxel incorporated into mixed micelles based on<br />

poly(ethylene glycol)–lipid conjugate and positively charged lipids, J. Drug Target., 13, 73, 2005.<br />

54. Weissig, V., Whiteman, K. R., and Torchilin, V. P., Accumulation of liposomal- and micellar-bound<br />

protein in solid tumor, Pharm. Res., 15, 1552, 1998.<br />

55. Lukyanov, A. N. et al., Polyethylene glycol-diacyllipid micelles demonstrate increased accumulation<br />

in subcutaneous tumors in mice, Pharm. Res., 19, 1424, 2002.<br />

56. Yuan, F. et al., Vascular permeability in a human tumor xenograft: Molecular size dependence and<br />

cutoff size, Cancer Res., 55, 3752, 1995.<br />

57. Hobbs, S. K. et al., Regulation of transport pathways in tumor vessels: Role of tumor type and<br />

microenvironment, Proc. Natl Acad. Sci. USA, 95, 4607, 1998.<br />

58. Monsky, W. L. et al., Augmentation of transvascular transport of macromolecules and nanoparticles<br />

in tumors using vascular endothelial growth factor, Cancer Res., 59, 4129, 1999.<br />

59. Helmlinger, G. et al., Interstitial pH and pO2 gradients in solid tumors in vivo: High-resolution<br />

measurements reveal a lack of correlation, Nat. Med., 3, 177, 1997.<br />

60. Tannock, I. T. and Rotin, D., Acid pH in tumors and its potential for therapeutic exploitation, Cancer<br />

Res., 49, 4373, 1989.<br />

61. Kwon, G. S. and Okano, T., Soluble self-assembled block copolymers for drug delivery, Pharm. Res.,<br />

16, 597, 1999.<br />

62. Cammas, S. et al., Thermorespensive polymer nanoparticles with a core-shell micelle structure as site<br />

specific drug carriers, J. Control. Rel., 48, 157, 1997.<br />

63. Chung, J. E. et al., Effect of molecular architecture of hydrophobically modified poly(N-isopropylacrylamide)<br />

on the formation of thermoresponsive core-shell micellar drug carriers, J. Control. Rel.,<br />

53, 119, 1998.<br />

64. Kohori, F. et al., Preparation and characterization of thermally responsive block copolymer micelles<br />

comprising poly(N-isopropylacrylamide-b-DL-lactide), J. Control. Rel., 55, 87, 1998.<br />

65. Meyer, O., Papahadjopoulos, D., and Leroux, J. C., Copolymers of N-isopropylacrylamide can trigger<br />

pH sensitivity to stable liposomes, FEBS Lett., 41, 61, 1998.<br />

66. Le Garrec, D. et al., Optimizing pH-responsive polymeric micelles for drug delivery in a cancer<br />

photodynamic therapy model, J Drug Target., 10, 429, 2002.<br />

67. Yoo, H. S., Lee, E. A., and Park, T. G., Doxorubicin-conjugated biodegradable polymeric micelles<br />

having acid-cleavable linkages, J Control. Rel., 82, 17, 2002.<br />

68. Bae, Y. et al., Multifunctional polymeric micelles with folate-mediated cancer cell targeting and pHtriggered<br />

drug releasing properties for active intracellular drug delivery, Mol. Biosyst., 1, 242–250,<br />

2005.<br />

69. Yokoyama, M. et al., Stabilization of disulfide linkage in drug-polymer-immunoglobulin conjugate <strong>by</strong><br />

microenvironmental control, Biochem. Biophys. Res. Commun., 164, 1234–1239, 1989.<br />

70. Nagasaki, Y. et al., Sugar-installed block copolymer micelles: Their preparation and specific<br />

interaction with lectin molecules, Biomacromolecules, 2, 1067, 2001.<br />

71. Vinogradov, S. et al., Polyion complex micelles with protein-modified corona for receptor-mediated<br />

delivery of oligonucleotides into cells, Bioconjug. Chem., 10, 851, 1999.<br />

72. Leamon, C. P., Weigl, D., and Hendren, R. W., Folate copolymer-mediated transfection of cultured<br />

cells, Bioconjug. Chem., 10, 947, 1999.<br />

73. Leamon, C. P. and Low, P. S., Folate-mediated targeting: From diagnostics to drug and gene delivery,<br />

Drug Discov. Today, 6, 44, 2001.<br />

74. Lee, E. S., Na, K., and Bae, Y. H., Polymeric micelle for tumor pH and folate-mediated targeting, J<br />

Control. Rel., 91, 103, 2003.<br />

75. Torchilin, V. P. et al., Immunomicelles: Targeted pharmaceutical carriers for poorly soluble drugs,<br />

Proc. Natl Acad. Sci. USA, 100, 6039, 2003.<br />

76. Torchilin, V. P. et al., p-Nitrophenylcarbonyl–PEG–PE–liposomes: Fast and simple attachment of<br />

specific ligands, including monoclonal antibodies, to distal ends of PEG chains via p-nitrophenylcarbonyl<br />

groups, Biochim. Biophys. Acta, 1511, 397, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


420<br />

Nanotechnology for Cancer Therapy<br />

77. Gao, Z. et al., PEG–PE/phosphatidylcholine mixed immunomicelles specifically deliver encapsulated<br />

taxol to tumor cells of different origin and promote their efficient killing, J. Drug Target., 11, 87, 2003.<br />

78. Iakoubov, L. Z. and Torchilin, V. P., A novel class of antitumor antibodies: Nucleosome-restricted<br />

antinuclear autoantibodies (ANA) from healthy aged nonautoimmune mice, Oncol. Res., 9, 439, 1997.<br />

79. Torchilin, V. P., Iakoubov, L. Z., and Estrov, Z., Therapeutic potential of antinuclear autoantibodies in<br />

cancer, Cancer Ther., 1, 179, 2003.<br />

80. Park, J. W. et al., Tumor targeting using anti-her2 immunoliposomes, J. Control. Rel., 74, 95, 2001.<br />

81. Sharma, A., Conway, W. D., and Straubinger, R. M., Reversed-phase high-performance liquid chromatographic<br />

determination of taxol in mouse plasma, J. Chromatogr. B. Biomed. Appl., 655, 315,<br />

1994.<br />

82. Ro<strong>by</strong>, A., Erdogan, S., and Torchilin, V. P., Solubilization of poorly soluble PDT agent, mesatetraphenylporphin,<br />

in plain or immunotargeted PEG–PE micelles results in dramatically improved<br />

cancer cell killing in vitro. Eur. J. Pharm. Biopharm., 62(3):235–240,<strong>2006</strong>.<br />

83. Mu, L., Elbayoumi, T. A., and Torchilin, V. P., Mixed micelles made of poly(ethylene glycol)–<br />

phosphatidylethanolamine conjugate and d-alpha-tocopheryl polyethylene glycol 1000 succinate as<br />

pharmaceutical nanocarriers for camptothecin, Int. J. Pharm., 306(1–2): 142–149, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


21<br />

CONTENTS<br />

Combined Cancer Therapy <strong>by</strong><br />

Micellar-Encapsulated Drugs<br />

and Ultrasound<br />

Natalya Rapoport<br />

21.1 Introduction ....................................................................................................................... 421<br />

21.2 Polymeric Micelles as Drug Carriers................................................................................ 422<br />

21.2.1 Advantages and Shortcomings of Polymeric Micelles as Drug Carriers .......... 422<br />

21.2.2 Polymeric Micelles in Clinical Trials ................................................................. 423<br />

21.3 Ultrasound in Drug Delivery ............................................................................................ 425<br />

21.3.1 Introduction to Ultrasound in Medicine ............................................................. 425<br />

21.3.2 Biological Effects of Ultrasound......................................................................... 425<br />

21.3.2.1 Cellular Level ..................................................................................... 425<br />

21.3.2.2 Systemic Level ................................................................................... 427<br />

21.3.3 Ultrasound-Induced Drug Release from Polymeric Micelles ............................ 428<br />

21.3.4 Ultrasound-Enhanced Intracellular Drug Uptake ............................................... 428<br />

21.4 In Vivo Evaluation of the Drug/Micelle/Ultrasound Tumor-Targeting Modality........... 430<br />

21.4.1 Ultrasound-Induced Drug Release from Micelles In Vivo ................................ 430<br />

21.4.2 Biodistribution of Polymeric Micelles and Micellar-Encapsulated Drugs ........ 430<br />

21.4.3 Effect of the Time of Ultrasound Application.................................................... 432<br />

21.4.4 Inhibition of Tumor Growth Using Micelle/Ultrasound Drug Delivery ........... 432<br />

21.4.4.1 Drug-Sensitive Tumors ...................................................................... 432<br />

21.4.4.2 Drug Resistant Tumors ...................................................................... 433<br />

21.4.4.3 Poorly Vascularized Tumors.............................................................. 435<br />

21.4.5 Mechanisms of Micelle/Ultrasound Bioeffects .................................................. 436<br />

21.5 Combining Ultrasound Imaging and Treatment ............................................................... 437<br />

21.5.1 Dual-Modality Contrast Agents/Drug Carrier Systems...................................... 437<br />

21.6 Conclusions ....................................................................................................................... 438<br />

References..................................................................................................................................... 438<br />

21.1 INTRODUCTION<br />

Chemotherapy is often complicated <strong>by</strong> toxic side effects of anti-cancer drugs. Also, effective<br />

chemotherapy regimens are frequently hindered <strong>by</strong> the resistance of tumor cells to one or more<br />

drugs [cross-resistance or multidrug resistance (MDR)]. The MDR is either inherent or acquired in<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

421


422<br />

the process of chemotherapy. Developing new tumor-localized chemotherapeutic modalities for<br />

treatment of drug-sensitive and MDR tumors is a major goal of current research in academia and<br />

industry. Towards this end, nanoparticle-based drug delivery combined with a localized drug<br />

release from carrier in the tumor volume is a promising approach to targeted chemotherapy.<br />

This chapter reviews the state of the art in drug targeting to tumors using polymeric micelles as<br />

drug carriers and tumor-localized ultrasound as a means to trigger drug release from micelles in the<br />

tumor volume. Ultrasound enhances the intracellular drug uptake and sensitizes MDR tumors to the<br />

action of conventional drugs.<br />

21.2 POLYMERIC MICELLES AS DRUG CARRIERS<br />

Nanotechnology for Cancer Therapy<br />

The polymeric micelles are formed <strong>by</strong> hydrophobic–hydrophilic block copolymers at concentrations<br />

above the critical micelle concentration (CMC). Below the CMC, block copolymer<br />

molecules exist in a solution in the form of individual molecules (unimers). At concentrations<br />

above the CMC, copolymer molecules self-assemble into dense micelles with hydrophobic cores<br />

and hydrophilic corona. The amphiphilic character of block copolymer micelles, their size<br />

(approximately 10–150 nm), and surface properties provide for a high drug loading capacity and<br />

long circulation time in the vascular system. This makes them attractive drug carriers. 1,2<br />

21.2.1 ADVANTAGES AND SHORTCOMINGS OF POLYMERIC MICELLES AS DRUG CARRIERS<br />

Drug encapsulation in micelles decreases systemic concentration of drug and diminishes intracellular<br />

drug uptake <strong>by</strong> normal cells, thus reducing unwanted drug interactions with healthy tissues.<br />

The important advantage offered <strong>by</strong> polymeric micelles is a so-called “enhanced penetration and<br />

retention (EPR) effect” 3 that provides for a selective accumulation of micellar-encapsulated drugs<br />

in tumor interstitium; in turn, this offers prospects for controlled drug release in the tumor volume.<br />

The simplicity of micelle formation <strong>by</strong> self-assembly of amphiphilic block copolymer<br />

molecules and drug encapsulation <strong>by</strong> physical mixing rather than chemical conjugation are extremely<br />

attractive features of polymeric micelles.<br />

Another important advantage offered <strong>by</strong> polymeric micelles is related to solubilization of drugs<br />

with poor aqueous solubility. Most anti-cancer drugs are lipophilic; thus, they have low aqueous<br />

solubility; conventional solubilizing agents, currently in use for the formulation of low-solubility<br />

drugs, are usually very toxic. Use of polymeric micelles as solubilizing agents results in dramatically<br />

increased aqueous solubility, and substantially decreased systemic toxicity of clinical<br />

formulations of paclitaxel (PTX), doxorubicin (DOX), and many other anti-cancer drugs. 4–18<br />

Various polymeric micellar systems have been designed to optimize important parameters of<br />

drug performance, such as aqueous solubility, on-demand release, and biological distribution. In<br />

this context, the most thoroughly studied are the micelles that comprise of poly(ethylene oxide)<br />

(PEO) as a hydrophilic block (or blocks). Examples of these include PEO-b-poly(propylene oxide)b-PEO<br />

copolymers (Pluronic w ), PEO-b-polyesters, and PEO-b-poly(amino acid)s.<br />

The application of biodegradable, pH-sensitive micelles like those of poly(ethylene glycol)-copoly(L-lactide)<br />

(PEG–PLLA) or poly(ethylene glycol)-co-poly(caprolactone) (PEG–PCL), may be<br />

especially attractive. If internalized <strong>by</strong> tumor cells via endocytosis, the micelles end up in the acidic<br />

environment of endosomes and lysosomes, where their accelerated degradation will release the<br />

encapsulated drug.<br />

In vitro drug release from micellar systems is usually assessed either <strong>by</strong> dialysis through<br />

semipermeable membranes or <strong>by</strong> gel-permeation chromatography. Unfortunately, the information<br />

provided <strong>by</strong> these techniques is not directly pertinent to the in vivo behavior of drug-loaded<br />

polymeric micelles. The systemic injection of micellar formulations is associated with substantial<br />

dilutions and sink conditions. Because micelle formation is thermodynamically driven, when<br />

diluted below the CMC, micelles dissociate into individual molecules (unimers), thus releasing<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 423<br />

their drug load. To prevent premature micelle dissociation and drug release upon injection, micellar<br />

systems should have low CMC, glassy cores, or cross-linked cores. This may be achieved <strong>by</strong><br />

induction of strong hydrophobic interactions or hydrogen bonds in the micelle cores. 19–21<br />

21.2.2 POLYMERIC MICELLES IN CLINICAL TRIALS<br />

Many micellar drug delivery systems showed promising results in in vitro and animal experiments,<br />

but showed disappointing behavior in clinical trials.<br />

The first clinical trials of micellar drug delivery systems began in this decade. Physically<br />

encapsulated DOX in PEO5000-b-P(Asp)4000-DOX micelles (NK911 formulation) entered clinical<br />

trials in Japan in 2001. 22 In animal studies, NK911 demonstrated about 30-fold higher area-underthe-curve<br />

(AUC) in plasma, higher tumor drug levels, lower toxicity, and higher therapeutic<br />

activity. However, in clinical trials, NK911 showed the same spectrum of side effects as free<br />

DOX, despite somewhat more favorable pharmacokinetic parameters. The discrepancy between<br />

the expected and observed clinical results was most likely caused <strong>by</strong> a premature drug release from<br />

micelles due to a very strong dilution of infused micellar solutions in the human circulatory system<br />

(volume: w5 L) when compared to that of mice (volume: w2 mL). A drop-like infusion of micellar<br />

drug delivery systems appears to have no prospect due to the dilution problem. Only bolus injections<br />

of reasonably stable micellar systems could provide for micelle preservation and drug<br />

retention in circulation.<br />

Aqueous Pluronic w solutions are another polymeric system suggested for DOX delivery,<br />

especially for drug resistant cancers. Pluronic w copolymers in unimeric concentrations were<br />

shown to hyper-sensitize drug resistance carcinoma cells and enhance response of resistant<br />

tumors to chemotherapeutic agents. 23–31 Thereasonbehindthisphenomenonisrelatedtoan<br />

increased drug uptake caused <strong>by</strong> deactivation of drug efflux pumps; the latter process is associated<br />

with Pluronic w -induced energy depletion in resistant tumor cells. 32 In addition, favorable changes<br />

in the intracellular trafficking of DOX occur in the MDR cells under Pluronic w action. 30,33,34<br />

Pluronic w formulation of DOX, known as SP1049C, was thoroughly investigated in animal<br />

studies <strong>by</strong> Alakhov et al. 35 Although this formulation is probably partly micellar, it is not expected<br />

to retain micellar structure upon intravenous injections. In vivo, DOX and Pluronic w unimers<br />

mostlikely move along independent routes. Not unexpectedly, the toxic level of DOX in<br />

SP1049C formulation was shown to be similar to that of free DOX. The results of pharmacokinetic<br />

and biodistribution experiments on SP1049C in Lewis lung carcinoma bearing C57BL/6 mice<br />

showed a very moderate increase in plasma AUC. In anti-tumor efficacy studies using hematological<br />

and solid animal tumor models, SP1049C was shown to be more effective than free DOX at<br />

an equal dose of 5 mg/kg. 30<br />

SP1049C entered preclinical and clinical trials in Canada in 1999. 36 The toxicity profiles of free<br />

DOX and Pluronic w formulation were identical, except for the histopathological changes in skin<br />

and thymus, that were less pronounced with SP1049C. In 2004, Danson et al. reported on the results<br />

of phase-I clinical trials of SP1049C. 36 In this study, SP1049C was given to patients with histologically<br />

proven cancer refractory to conventional treatments. The dose-limiting toxicity (DLT) in<br />

humans, i.e., myelosuppression, was reached at 90 mg/m 2 of SP1049C; therefore, a dose of 70 mg/<br />

m 2 was recommended for future trials. The pharmacokinetic profile of SP1049C was similar to that<br />

of a clinical DOX formulation, except for a slower terminal clearance stage. SP1049C induced<br />

temporal partial response in several patients with advanced solid tumors. The results were<br />

considered promising, and the phase-II trials were initiated; their first results were recently reported<br />

<strong>by</strong> Valle et al. 37 In phase-II trials, SP1049C was evaluated in patients with less severe forms of<br />

cancer. Patients received 75 mg/m 2 of SP1049C (30 min intravenous infusion), four times a week,<br />

for up to six cycles. The authors reported partial responses in some patients after 4–6 treatment<br />

cycles. However, data showed appearance of hematological and nonhematological (especially<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


424<br />

Nanotechnology for Cancer Therapy<br />

cardiac) toxicities, presumably because DOX was not encapsulated and, therefore, not prevented<br />

from the attack on the heart muscle.<br />

PTX is another potential drug candidate for the delivery in polymeric micelles. PTX is a highly<br />

hydrophobic molecule; for clinical use, PTX is solubilized with the aid of a significant amount of<br />

Cremophor EL and ethyl alcohol. 38 These formulations (Taxol w or PTX Injections) cause significant<br />

systemic toxicity. Considerable efforts of various groups have been dedicated to developing<br />

micellar formulations of PTX. 4–6,9,39 However, except for the Japanese formulation NK105 and<br />

Genexol w (see below), no other formulations have progressed to clinical trials due to a rapid loss of<br />

drug from micellar carriers, resulting in the absence of advantages over clinical formulations.<br />

In 2001, Kim et al. reported on PTX-loaded PEO-b-poly(D,L-lactide) (PEO2000-b-PDLLA1750)<br />

micelles (Genexol w -PM) prepared <strong>by</strong> a solvent evaporation technique. 39 The maximum tolerated<br />

doses (MTDs) of Genexol w -PM and Taxol w for the intravenous (i.v.) administration in nude mice<br />

were 60 and 20 mg/kg, respectively. This allowed a threefold higher administration dose of drug. In<br />

animal studies, Genexol w -PM showed a better anti-tumor effect in ovarian and breast cancer mice<br />

models than clinical Taxol w .<br />

However, despite an increase in the administered dose, plasma AUC was 20-fold lower for the<br />

micellar formulation, which was attributed to a fast release of drug from the micelles. Another<br />

factor involved in the fast elimination of micellar PTX from the circulation could be related to the<br />

uptake <strong>by</strong> the cells of the reticulo-endothelial system (RES). In all organs except plasma, PTX<br />

biodistribution was comparable for a clinical formulation or Genexol-PM. In phase-I human trials,<br />

micellar-encapsulated PTX allowed administration of a higher drug dose, but resulted in a lower<br />

plasma AUC and a shorter PTX half life. 40<br />

Another type of polymeric micelles suggested for PTX delivery comprise PDLLA as a coreforming<br />

block and poly(N-vinylpyrrolidone) (PVP) as a shell-forming block. 41 The in vivo evaluation<br />

of this system showed results similar to those for Genexol w -PM in regards to a reduced<br />

plasma AUC. 42 For the micellar formulation, the AUC was also reduced in liver, kidney, spleen,<br />

and heart, but was not significantly changed in the tumor. For the polymeric micellar formulation,<br />

MTD was not reached even at 100 mg/kg, whereas the MTD of Taxol w was 20 mg/kg. This allowed<br />

a threefold increase of the therapeutic dose of PTX, resulting in a significant increase of the<br />

anti-tumor activity of the polymeric system compared to Taxol w .<br />

Polymeric micellar PTX formulation NK105 reported <strong>by</strong> Hamaguchi et al. 43 comprises<br />

micelle-forming block copolymer poly(ethelene oxide)-b-poly(4-phenyl-1-butanoate)-L-apartamide<br />

(PEO-b-PPBA). This formulation demonstrated stealth and targeting properties. NK105<br />

manifested a 86-fold increase in PTX plasma AUC compared to Taxol w , accompanied <strong>by</strong> a<br />

25-fold increase for drug AUC in tumor, and a higher anti-tumor activity in C-26 tumor-bearing<br />

mice. At a PTX dose of 100 mg/kg, tumors were completely resolved after a single administration<br />

of NK105. This formulation is currently in phase-I clinical trials in Japan.<br />

In summary, only a few polymeric micellar formulations have demonstrated success in clinical<br />

trials. Polymeric micelles as drug carriers are either too stable, therefore incapable of providing<br />

adequate drug release at the tumor site, or too unstable, thus prematurely releasing their<br />

drug payload.<br />

To overcome these complications, we have developed Pluronic w -based micellar systems that<br />

withstand the destabilizing effects of the biological environment and effectively target tumors while<br />

retaining their drug load. 44 Various routes of micelle stabilization were tested for Pluronic<br />

micelles; 19 the optimal stability, low toxicity, and long circulation time of micelle-encapsulated<br />

DOX was achieved <strong>by</strong> forming mixed micelles of Pluronic w and PEO-diacyl phospholipids 44 or <strong>by</strong><br />

introducing a small concentration of oil in the micelle core to increase core hydrophobicity. 19 The<br />

on-demand release of DOX from micelles and an effective intracellular drug uptake <strong>by</strong> the tumor<br />

cells was achieved <strong>by</strong> local ultrasound treatment of the tumor at a particular time after the drug<br />

injection. 44<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 425<br />

Important advantages of ultrasound are that it is noninvasive, can penetrate deep into the<br />

interior of the body, and can be focused and carefully controlled. Therapeutic tumor treatment<br />

<strong>by</strong> ultrasound requires tumor imaging <strong>by</strong> diagnostic ultrasound prior to therapy. During the last<br />

decade, combining diagnostic and therapeutic ultrasound in the same instrument has attracted ever<br />

growing attention. Developing dual-modality systems that combine properties of ultrasound<br />

contrast agents and drug carriers is a timely task.<br />

21.3 ULTRASOUND IN DRUG DELIVERY<br />

21.3.1 INTRODUCTION TO ULTRASOUND IN MEDICINE<br />

Ultrasound consists of pressure waves generated <strong>by</strong> piezoelectric transducers, with frequencies at or<br />

greater than 20 kHz. Like optical or audio waves, ultrasonic waves can be focused, reflected,<br />

refracted, and propagated through a medium, thus allowing the waves to be directed to and<br />

focused on a particular volume of tissue in a spatially and temporally controlled manner. Ultrasound<br />

technology allows for a high degree of spatial and dynamic control due to a favorable range<br />

of energy penetration characteristics in soft tissue and the ability to shape energy<br />

deposition patterns.<br />

In medicine, ultrasound is used as either a diagnostic or a therapeutic modality. The main<br />

advantage of ultrasound is its noninvasive nature; the transducer is placed in contact with a waterbased<br />

gel or water layer on the skin, and no insertion or surgery is required.<br />

In current therapeutic applications, ultrasound is used mostly for tissue thermoablation.<br />

Progress in imaging techniques has allowed ablating large tumor masses <strong>by</strong> high intensity<br />

focused ultrasound (HIFU). HIFU instruments noninvasively deliver high intensity ultrasound<br />

energy to predetermined volumes of the body.<br />

However, the ablative techniques require delivering high acoustic energy to a well delineated<br />

volume in the body over a substantial time period (up to several hours), which is associated with a<br />

potential risk of ablating surrounding tissues due to a patient’s breathing or motion. Motion<br />

correction is a serious concern in HIFU applications. In addition, some locations in the body are<br />

screened <strong>by</strong> bones and not accessible for HIFU treatment.<br />

In our lab, we have been developing a different mode of application of therapeutic ultrasound.<br />

In this modality, tumor cell killing is based on the ultrasound-enhanced cytotoxic action of tumortargeted<br />

chemotherapeutic drugs rather than the mechanical or thermal action of ultrasound. 45–56<br />

This technology requires order of magnitude lower ultrasound energy, reducing the risk of coagulative<br />

necrosis of healthy tissues associated with HIFU.<br />

21.3.2 BIOLOGICAL EFFECTS OF ULTRASOUND<br />

21.3.2.1 Cellular Level<br />

The effect of ultrasound on the cellular level is usually attributed to formation of plasma membrane<br />

defects (sonoporation) induced <strong>by</strong> cavitating microbubbles; the formation and collapse of microbubbles<br />

is called inertial cavitation. This process has been investigated using a voltage clamp<br />

technique. 57–59 It was shown that sonoporation was substantially enhanced <strong>by</strong> ultrasound contrast<br />

agents that nucleate inertial cavitation. 60 The degree of membrane damage increased with<br />

increasing duty cycle of pulsed ultrasound. 59 Some membrane defects are resealed when ultrasound<br />

is turned off (this is called transient sonoporation); a key role in membrane resealing is played <strong>by</strong><br />

Ca 2C . 58<br />

In our experiments, ultrasound bioeffects on the cellular level were compared for the drugsensitive<br />

and multidrug resistant (MDR) A2780 ovarian carcinoma cells in the presence or absence<br />

of DOX. 61 Focused 1.1-MHz ultrasound was applied at varying peak negative pressures and DOX<br />

concentrations. Cell viability, a degree of membrane damage, DOX uptake, and cell proliferation<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


426<br />

were measured. Cell incubation with the nucleic acid stain Sytox allowed discriminating between<br />

the cells experiencing various degrees of plasma membrane damage. Fluorescence of Sytox is<br />

negligible in the extracellular environment; when Sytox enters the cells and intercalates the<br />

DNA, its fluorescence increases <strong>by</strong> several orders of magnitude. For the present application it is<br />

important that Sytox does not penetrate into the cells with intact membranes; however, when cell<br />

membranes are damaged, Sytox enters the cells and intercalates the DNA, which results in a<br />

dramatic fluorescence increase.<br />

Major results of our experiments on ultrasound bioeffects are listed below.<br />

† Under the action of ultrasound, cells with various degrees of membrane damage, intermediate<br />

between viable and dead cells, were generated.<br />

† At the same delivered ultrasound energy, higher ultrasound pressure caused more severe<br />

membrane damage.<br />

† A proliferation rate of sonicated multidrug resistant ovarian carcinoma A2780/AD cells<br />

with compromised membranes was significantly lower than the proliferation rate of the<br />

intact cells (Figure 21.1).<br />

† The MDR A2780/AD cells manifested a higher susceptibility to ultrasound than parental<br />

drug-sensitive A2780 cells (Figure 21.2).<br />

† Introduction of DOX in the incubation medium increased cell susceptibility to the<br />

ultrasound action. For the same ultrasound energy, the cells sonicated in the presence<br />

of DOX manifested significantly higher membrane damage than the cells without DOX<br />

(Figure 21.3). For the same initial DOX concentration in the incubation medium, the<br />

effect of DOX was stronger at 378C than at room temperature suggesting that an<br />

increased cell susceptibility to ultrasound was caused <strong>by</strong> DOX internalized <strong>by</strong> the<br />

cells. We hypothesize that the effect of DOX may be associated with a destabilizing<br />

action on the DNA molecule at the intercalation sites.<br />

Very interesting data on the effect of ultrasound on the gene expression in multidrug resistant<br />

Hep2D/ADR hepatic carcinoma cells were recently obtained in Chongqing Medical University <strong>by</strong><br />

Shao and Wu (Wu, F., personal communication). These researchers observed a noticeable suppression<br />

of the MDR1 gene expression in the cells sonicated <strong>by</strong> 0.8-MHz ultrasound; the BCL2/BAX<br />

expression ratio was also substantially reduced, resulting in an enhanced apoptosis. The last effect<br />

was significantly stronger in the MDR compared to parental cells.<br />

Percent survival/growth<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Immediate viability<br />

Nanotechnology for Cancer Therapy<br />

no DOX<br />

[DOX]= 1 ug/ml<br />

[DOX]= 20 ug/ml<br />

Proliferation<br />

FIGURE 21.1 Immediate cell viability (left group of bars) and proliferation at 378C for 72 h (right group<br />

of bars) for the A2780/AD cells sonicated (0.55 MPa, 15 s) in the presence or absence of various concentrations<br />

of DOX. (Adapted from Kamaev, P., Christensen, D. A., and Rapoport, N., unpublished results.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 427<br />

Viability, %<br />

These clinically important findings can be used for effective therapy of multidrugresistant<br />

tumors.<br />

21.3.2.2 Systemic Level<br />

In living organisms, the bioeffects of ultrasound are not limited to sonoporation. On a systemic<br />

level, a very important role may be played <strong>by</strong> enhanced extravasation. Because of the adverse<br />

effects that this may have in diagnostic ultrasound, extensive work has been done on studying<br />

extravasation at various locations in the body and from various blood vessels, especially in<br />

the process of heart imaging. Extravasation was substantially enhanced <strong>by</strong> ultrasound contrast<br />

agents. Enhanced leakage of the die and erythrocytes from heart vessels was observed in several<br />

studies. 62–64 Erythrocyte extravasation under the action of contrast-enhanced ultrasound was also<br />

observed in the kidney; it was shown that the effect of pulsed ultrasound was stronger than that of<br />

continuous wave (CW) ultrasound. 65<br />

Although enhanced extravasation is an undesirable effect in ultrasound imaging, it may play a<br />

very positive role in drug delivery, especially for the local delivery of macromolecular drugs or<br />

Viabilty, %<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0<br />

A2780<br />

500<br />

Energy exposure, J/cm 2<br />

1000<br />

A2780<br />

A2780/AD<br />

FIGURE 21.2 Immediate cell viability after ultrasound exposure decreases with increasing acoustic energy;<br />

cells were irradiated in suspensions <strong>by</strong> focused 1.1-MHz ultrasound. The MDR cells manifest higher susceptibility<br />

to the ultrasound action than their wild-type counterparts. (Adapted from Kamaev, P., and Rapoport, N.,<br />

Therapeutic Ultrasound: 5th International Symposium on Therapeutic Ultrasound, Clement, G. T., McDannold,<br />

N. J., and Hynynen, K., Eds., American Institute of Physics, Melville, NY, pp. 543–547, <strong>2006</strong>. With permission.)<br />

A2780/AD<br />

no DOX<br />

DOX<br />

FIGURE 21.3 Effect of DOX on the immediate viability of drug-sensitive and MDR ovarian carcinoma cells<br />

sonicated in suspensions <strong>by</strong> focused ultrasound; DOX concentration in the media is 10 mg/mL for the MDR<br />

cells and 20 mg/mL for the sensitive cells. Ultrasound conditions: frequency 1.1 MHz, pressure 0.55 MPa,<br />

duration 15 s, duty cycle 33%, room temperature. (Adapted from Kamaev, P., and Rapoport, N., Therapeutic<br />

Ultrasound: 5th International Symposium on Therapeutic Ultrasound, Clement, G. T., McDannold, N. J., and<br />

Hynynen, K., Eds., American Institute of Physics, Melville, NY, pp. 543–547, <strong>2006</strong>. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

1500


428<br />

drug-loaded nanoparticles. It was shown that ultrasound enhanced the extravasation of the macromolecular<br />

MRI contrast agent. 66<br />

21.3.3 ULTRASOUND-INDUCED DRUG RELEASE FROM POLYMERIC MICELLES<br />

Ultrasound was shown to release DOX from the unstabilized and stabilized Pluronic w micelles. 46,55<br />

A degree of drug release depended on a number of ultrasound parameters—frequency, power<br />

density, pulse length, and inter-pulse intervals. 46 Examples of drug release profiles under CW or<br />

pulsed ultrasound are presented in Figure 21.4. Measurements were based on a decrease of DOX<br />

fluorescence when drug was transferred from the hydrophobic environment of micelle cores into the<br />

aqueous environment. Measurements were performed in situ in the absence of cells.<br />

As indicated <strong>by</strong> fluorescence profiles, drug release under the action of ultrasound pulses was<br />

reversible and drug re-encapsulation proceeded during the inter-pulse intervals. Pulsed ultrasound<br />

induced alternating cycles of micelle perturbation and restoration. Kinetic parameters of the drug<br />

release and re-encapsulation indicated that both processes were controlled <strong>by</strong> motion of micelleforming<br />

macromolecules rather than diffusion of relatively small drug molecules. 55<br />

Low-frequency ultrasound was more effective in drug release from micelles than highfrequency<br />

ultrasound (Figure 21.5). Low-frequency ultrasound can penetrate deeper into the<br />

tissue than high-frequency ultrasound, but it does not allow sharp focusing. 50 Therefore, in vivo,<br />

high-frequency ultrasound could be used for irradiating small and relatively superficial tumors,<br />

while low-frequency ultrasound could be used for larger and deeper located tumors.<br />

21.3.4 ULTRASOUND-ENHANCED INTRACELLULAR DRUG UPTAKE<br />

Without ultrasound, drug encapsulation in polymer micelles results in a significantly reduced intracellular<br />

uptake. The degree of drug shielding depends on drug/micelle interaction, and is higher for<br />

Fluorescence intensity, Arb, Uniits<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

Continuous wave<br />

Pulsed<br />

0.5s "on" : 0.5s "off"<br />

1s "on" : 1s "off"<br />

1s "on" : 3s "off"<br />

0 20 40 60 80 100 120 140 160<br />

Time (s)<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 21.4 Doxorubicin (DOX) release profiles from 10% Pluronic micelles under continuous wave (CW)<br />

and pulsed 20 kHz ultrasound at a power density of 0.058 W/cm 2 at various durations of ultrasound pulses and<br />

inter-pulse intervals indicated in the figure. Measurements are based on the decrease of fluorescence intensity<br />

when DOX is transferred from the hydrophobic environment of micelles cores into the aqueous environment.<br />

(Adapted from Husseni, G. A. et al., J. Control. Release, 69, 43–52, 2000. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

180


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 429<br />

Degree of drug release, %<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

67 kHz<br />

1 MHz<br />

0<br />

0 2 4<br />

Power density, W/cm<br />

6 8<br />

2<br />

FIGURE 21.5 Effect of ultrasound frequency and power density on DOX release from 10% Pluronic P-105<br />

micelles. (Adapted from Marin, A. et al., J. Control. Release, 84, 39–47, 2002.)<br />

more lipophilic drugs. As an example, DOX uptake <strong>by</strong> ovarian carcinoma A2780 cells dropped <strong>by</strong><br />

50% after encapsulation in 10% Pluronic w micelles. For a more hydrophobic analog of DOX called<br />

Ruboxyl, micellar encapsulation resulted in a decrease of the intracellular uptake <strong>by</strong> 66%.<br />

Drug shielding <strong>by</strong> micelles reduces systemic concentrations of toxic drugs, diminishing side<br />

effects. As an example, DOX encapsulation in PEO-b-poly(a-benzyl L-aspartate) (PEO–PBLA)<br />

micelles allowed administering high concentrations of DOX to tumor-bearing mice without<br />

toxicsideeffects. 67 The same results were manifested <strong>by</strong> micelle-encapsulated PTX 39–43<br />

(see Section 21.2.2).<br />

Ultrasound substantially enhanced the intracellular drug uptake from/with polymeric<br />

micelles. 45,50,56 Typical examples are shown in Figure 21.6.<br />

Based on the results presented above, a new modality of tumor-targeted chemotherapy has been<br />

suggested. According to this technology, micelle-encapsulated drugs are injected intravenously to<br />

tumor-bearing subjects. Drug-loaded micelles are long-circulating (Figure 21.7), which provides<br />

for their gradual accumulation in the tumor interstitium via the EPR effect (see below). At the time<br />

corresponding to a maximal micelle accumulation, ultrasound is applied locally to the tumor. Under<br />

the ultrasound action, drug is released from micelles in the tumor volume; the intracellular uptake<br />

of drug <strong>by</strong> tumor cells is enhanced due to ultrasound-induced plasma membrane perturbation.<br />

Important advantages of this technique are its noninvasive character, on-demand drug release,<br />

and a prospect for sensitization of drug-resistant tumors.<br />

Counts<br />

10 0<br />

600<br />

480<br />

360<br />

240<br />

120<br />

0<br />

1<br />

2<br />

10 1<br />

3<br />

4<br />

10<br />

Fluorescence, arb.u.<br />

2<br />

FIGURE 21.6 Fluorescence histogram of A2780 cells. (1) Control; (2)–(4): cells incubated for 30 min. with<br />

20 mg/mL DOX encapsulated in PEG–PLLA micelles. (2) No ultrasound; (3) sonication for 30 s <strong>by</strong> CW<br />

3-MHz ultrasound at 2.0 W/cm 2 ; (4) sonication for 30 s <strong>by</strong> 1-MHz ultrasound at 3.4 W/cm 2 and 33% duty<br />

cycle. (From Gao, Z. and Rapoport, N., unpublished results.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

10 3<br />

10 4


430<br />

DOX Concentration in plasma, μg/ml<br />

10<br />

1<br />

0.1<br />

0.01<br />

0 4 8 12<br />

Time, h<br />

16 20 24<br />

In the following section, the results of the in vivo evaluation of the new therapeutic modality<br />

are discussed.<br />

21.4 IN VIVO EVALUATION OF THE DRUG/MICELLE/ULTRASOUND<br />

TUMOR-TARGETING MODALITY<br />

21.4.1 ULTRASOUND-INDUCED DRUG RELEASE FROM MICELLES IN VIVO<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 21.7 Pharmacokinetics of DOX dissolved in PBS (squares) or encapsulated in stabilized Pluronic<br />

P-105 micelles (diamonds); mean values plus/minus standard deviations are presented, nZ4. (From Gao, Z.<br />

and Rapoport, N., unpublished results.)<br />

Drug release from micelles in vivo was confirmed in experiments described in studies <strong>by</strong> Gao,<br />

et al. 44 Ovarian carcinoma A2780 cells (2!10 6 ) were inoculated in peritoneal cavities of athymic<br />

(nu/nu) mice. The next day, mice were treated intraperitoneally <strong>by</strong> 3 mg/kg DOX that was either<br />

dissolved in PBS (designated “free DOX” hereafter) or encapsulated in 10% Pluronic w P-105<br />

micelles. Treatment was combined or not combined with the ultrasonic irradiation of the abdominal<br />

region of a mouse. Some groups of mice were treated <strong>by</strong> ultrasound without DOX. Tumor yields<br />

and animal survival rates were monitored. The results are presented in Table 21.1. 44<br />

Without DOX, no effect of ultrasound on the tumor staging was observed; the tumors were<br />

developed in 100% mice. Only a moderate effect of free DOX on the tumor development was<br />

observed without ultrasound, tumor yield being reduced <strong>by</strong> 30%. However, when DOX and ultrasound<br />

were combined, tumor yields were reduced <strong>by</strong> about 60% for both free and micellarencapsulated<br />

DOX, which was accompanied <strong>by</strong> a statistically significant increase of the average<br />

survival rates of corresponding animal groups. No differences were observed between the effects of<br />

free or micellar-encapsulated DOX, suggesting that ultrasound did, indeed, trigger DOX release<br />

from micelles in vivo.<br />

The absence of the ultrasound effect in the absence of DOX indicated that tumor cell death was<br />

associated with the ultrasound-enhanced cytotoxic action of DOX rather than mechanical or<br />

thermal action of ultrasound itself. 44<br />

21.4.2 BIODISTRIBUTION OF POLYMERIC MICELLES AND MICELLAR-ENCAPSULATED DRUGS<br />

The effect of ultrasound on biodistribution of polymeric micelles was studied using fluorescently<br />

labeled Pluronic w P-105 micelles. 47 The uptake of micelles <strong>by</strong> various organ cells was measured <strong>by</strong><br />

flow cytometry. A significant advantage of this technique is that it does not require organ<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 431<br />

TABLE 21.1<br />

Tumor Yields and Survival Rates of Mice Inoculated on Day 0 with i.p. Injections of A2780<br />

Ovarian Carcinoma Cells<br />

<strong>Group</strong><br />

Number of Animals<br />

in <strong>Group</strong><br />

Number of Internal<br />

Tumors Tumor Yield (%) Survival Rate (Days)<br />

PBS 17 17 100 40.6G8.2<br />

PBSCUS 5 5 100 39.6G6.5<br />

P-105 empty micelles 13 13 100 51.9G21.6<br />

P-105 empty<br />

micelles/US<br />

13 13 100 47.2G9.0<br />

DOX/PBS a<br />

10 7 70 61.5G10 b<br />

DOX/PBS/US a<br />

5 2 40 77.3G24<br />

DOX/micelle/US a<br />

11 4 36 91G27 b<br />

a<br />

The mice were treated on days 1, 4, 7, and 11; DOX at 3 mg/kg in various delivery systems was i.p. injected. One hour after<br />

the DOX injection, some groups of animals were sonicated in the abdominal region for 30 s <strong>by</strong> 1-MHz ultrasound at a power<br />

density of 1.2 W/cm 2 . By day 25, all control mice developed tumors. The last three groups marked <strong>by</strong> “a” were treated on<br />

days 1, 26, and 35.<br />

b pZ.01.<br />

Source: Gao, Z., Fain, H., and Rapoport, N., J. Control. Release, 102, 203–221, 2005.<br />

homogenization, thus providing for direct measurements of carrier or drug concentrations inside the<br />

cells, i.e., at the site of action.<br />

A 100-mL solution of fluorescently labeled stabilized Pluronic w micelles was injected intravenously<br />

through the tail vein to ovarian carcinoma bearing nu/nu mice. At various time points<br />

ranging from 2 to 12 h after the i.v. injection, mice were sonicated for 30 s <strong>by</strong> 1-MHz or 3-MHz<br />

ultrasound. Ten minutes after the ultrasound application, mice were sacrificed, tumor and other<br />

organs excised, dried with filter paper, weighed, and sliced in trypsin. The individual cells of<br />

various organs were fixed <strong>by</strong> 2.5% glutaraldehyde. Fluorescence of Pluronic w P-105 molecules<br />

internalized <strong>by</strong> the cells of various organs was measured <strong>by</strong> flow cytometry.<br />

Without ultrasound, a multimodal distribution of tumor cell fluorescence was observed for<br />

the intravenous injections of the fluorescently labeled, stabilized Pluronic w micelles. A significant<br />

fraction of the cells manifested a relatively low fluorescence (Figure 21.8, regular line). Fluorescence<br />

intensity of other organ cells was at the level of the autofluorescence of corresponding<br />

cells. These data indicated low efficiency of the intracellular uptake of intact Pluronic w micelles.<br />

A short (30-s ultrasound exposure) tumor sonication strongly enhanced the intracellular uptake of<br />

Pluronic w molecules <strong>by</strong> the tumor cells (Figure 21.8, bold line), suggesting that before sonication,<br />

micelles were effectively accumulated in the tumor interstitium but were not effectively internalized<br />

<strong>by</strong> the tumor cells. Tumor sonication not only substantially enhanced micelle internalization,<br />

but also resulted in a much more uniform fluorescence distribution over the tumor cell population,<br />

suggesting an enhanced micelle diffusion over the tumor volume. 47<br />

The biodistribution of micellar-encapsulated DOX closely followed that of a micellar carrier. 44<br />

For micellar-encapsulated DOX without ultrasound, DOX uptake <strong>by</strong> the tumor cells was relatively<br />

low. A local tumor sonication <strong>by</strong> 1-MHz ultrasound dramatically increased the intracellular uptake<br />

of drug <strong>by</strong> the tumor cells (Figure 21.9). A similar effect was observed for PEG–PBLA micelles<br />

(Figure 21.10). 44 Somewhat enhanced DOX uptake <strong>by</strong> the cells of other organs was observed for<br />

1-MHz ultrasound (data not shown). This was associated with a deep penetration of 1-MHz ultrasound<br />

in the mouse body, resulting in ultrasound scattering at various internal interfaces and/or<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


432<br />

Counts<br />

10 0<br />

150<br />

120<br />

90<br />

60<br />

30<br />

0<br />

reflection from the skin/air interface at the site opposite to the transducer. This unwanted effect was<br />

eliminated when tumor was irradiated <strong>by</strong> a more localized 3-MHz ultrasound (Figure 21.10).<br />

21.4.3 EFFECT OF THE TIME OF ULTRASOUND APPLICATION<br />

Measuring micelle accumulation in the tumor cells for various times of ultrasound application after<br />

drug injection enabled the selection of optimal ultrasound application times. For the stabilized<br />

Pluronic w micelles, the optimal time of ultrasound application was found to be between four and<br />

eight hours after injection. 44,47 The results were interpreted as follows. At shorter times, a large<br />

number of micelles were still circulating in the vasculature while a lower concentration of micelles<br />

was accumulated in the tumor interstitial space. At longer times, some clearance of micelles from<br />

the tumor volume could already have occurred. The optimal time of ultrasound application is<br />

expected to depend on the pharmacokinetics of a particular delivery system.<br />

21.4.4 INHIBITION OF TUMOR GROWTH USING MICELLE/ULTRASOUND DRUG DELIVERY<br />

21.4.4.1 Drug-Sensitive Tumors<br />

1<br />

10 1<br />

10<br />

Fluorescence, arb.u.<br />

2<br />

FIGURE 21.8 Fluorescence histogram of the tumor cells in (1) unsonicated, and (2) sonicated mouse; 5%<br />

stabilized Pluronic micelles comprising 0.1% fluorescently labeled Pluronic P-105 were injected intravenously<br />

to two subcutaneous A2780 tumor bearing mice; 4 h after the injection, the tumor of one mouse was sonicated<br />

for a total of 60 s <strong>by</strong> 1-MHz ultrasound at 3.4 W/cm 2 power density at 50% duty cycle (two 30-s sonications<br />

were performed, with a 30-s interval between the sonications for cooling the ultrasound probe); both mice were<br />

sacrificed 10 min after the completion of sonication. (Modified from Gao, Z., Fain, H., and Rapoport, N., Mol.<br />

Pharmaceutics, 1, 317–330, 2004.)<br />

Targeting of micellar-encapsulated drugs to tumor cells and uniform drug distribution in the tumor<br />

volume result in a significant suppression of the growth of subcutaneous ovarian carcinoma tumors<br />

Counts<br />

10 0<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1<br />

10 1<br />

2<br />

10<br />

Fluorescence, arb.u.<br />

2<br />

FIGURE 21.9 Fluorescence histograms of the tumor cells in (1) nonsonicated, and (2) sonicated mice injected<br />

i.v. with DOX (6 mg/kg) encapsulated in mixed micelles. Mice were sonicated for 30 s <strong>by</strong> 1-MHz CW<br />

ultrasound at a power density of 1.7 W/cm 2 applied 8 h after the drug injection; nonsonicated and sonicated<br />

mice were sacrificed 10 min later. (Modified from Gao, Z., Fain, H., and Rapoport, N., J. Control. Release,<br />

102, 203–221, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

10 3<br />

10 3<br />

2<br />

10 4<br />

10 4


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 433<br />

Counts<br />

(a)<br />

(b)<br />

(c)<br />

(d)<br />

10 0<br />

inoculated in nu/nu mice (Figure 21.11). 44 However, the success of the treatment depended<br />

dramatically on the stage of tumor progression at the start of the treatment. When the initial<br />

tumor volume at the start of the treatment exceeded 250 mm 3 , none of the treatment modalities<br />

was effective in suppressing tumor growth, which emphasizes a need for early tumor detection.<br />

21.4.4.2 Drug Resistant Tumors<br />

10 1<br />

1<br />

We have attained substantial success in treating multidrug resistant, poorly differentiated MCF-<br />

7/ADmt breast cancer tumors <strong>by</strong> micellar-encapsulated PTX combined with ultrasonic tumor irradiation.<br />

The MCF7/AD mt tumor cells expressed the MDR1 gene 25 and were highly resistant to<br />

treatment <strong>by</strong> clinical PTX formulation, Taxol w (PTX injection). However, tumor growth was<br />

substantially suppressed when animals were treated <strong>by</strong> intravenous injections of a micellar PTX<br />

formulation in conjunction with local tumor sonication (Figure 21.12) (the micellar formulation of<br />

PTX, Genexol w was kindly donated <strong>by</strong> Samyang Pharmaceuticals Inc. (Korea)).<br />

In vitro without ultrasound, the intracellular uptake of PTX from the micellar formulation was<br />

significantly lower than that with PTX injection, which can be advantageous when used in vivo for<br />

preventing unwanted drug interactions with healthy cells. Under ultrasound, drug uptake <strong>by</strong> the<br />

MCF7/AD mt cells from micellar PTX formulation was increased more than 20-fold, which resulted<br />

in the inhibition of cellular proliferation <strong>by</strong> nearly 90%.<br />

10 2<br />

Fluorescence,arb.u.<br />

2<br />

10 3<br />

Kidney<br />

Spleen<br />

FIGURE 21.10 Fluorescence histograms of (a) tumor, (b) kidney, (c) spleen, and (d) heart cells of (1)<br />

nonsonicated (regular lines), and (2) sonicated (bold lines) mice injected with DOX (6 mg/kg) encapsulated<br />

in PEG–PBLA micelles. Mice were sonicated for 30 s <strong>by</strong> 3-MHz CW ultrasound 12 h after the drug injection at<br />

a power density of 1.8 W/cm 2 . (Adapted from Gao, Z., Fain, H., and Rapoport, N., J. Control. Release, 102,<br />

203–221, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Heart<br />

10 4


434<br />

Tumor volume, mm 3<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

0 10 20 30 40<br />

Days after first treatment<br />

control<br />

DOX/PBS<br />

DOX/micelles<br />

DOX/micelles/US<br />

FIGURE 21.11 Growth curves of s.c. A2780 ovarian carcinoma tumors inoculated in female nu/nu mice;<br />

DOX (3 mg/kg) was either dissolved in PBS (positive control) or encapsulated in mixed micelles; drug<br />

injections were combined or not combined with tumor sonication. Treatments were initiated when the<br />

tumor volume reached 75–125 mm 3 ; three consecutive treatments were applied on days 1, 3, and 5. Ultrasound<br />

parameters: frequency, 1 MHz; power density, 3.4 W/cm 2 ; duty cycle 50%; duration, 30 s; time of application:<br />

4 h after the i.v. injection of the drug. (Adapted from Gao, Z., Fain, H., and Rapoport, N., J. Control. Release,<br />

102, 203–221, 2005.)<br />

The in vivo results are shown in Figure 21.12. As could be expected, without ultrasound,<br />

clinical PTX and Genexol w both manifested low efficacy; in contrast, injections of Genexol w<br />

combined with ultrasonic tumor irradiation resulted in a complete tumor resolution. Most importantly,<br />

the tumors treated <strong>by</strong> Genexol w /ultrasound did not metastasize, whereas tumors treated<br />

without ultrasound manifested a high rate of metastasizing into lymph nodes and/or spleen.<br />

Total Tumor Volume (mm 3 )<br />

550<br />

500<br />

Control<br />

Clinical paclitaxel<br />

Genexol<br />

450 Genexol + US<br />

400<br />

350<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

0 20 40<br />

Time (days)<br />

60 80<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 21.12 Growth curves of subcutaneous MCF7/ADmt tumors inoculated in female nu/nu mice; bolus<br />

injections of a clinical paclitaxel formulation Paclitaxel injections or a micellar formulation Genexol w<br />

provided <strong>by</strong> Samyang Pharmaceuticals, Inc. (Korea) were administered twice a week for two weeks intravenously<br />

through the tail vein at a dose rate of 13 mg/kg/day. (From Howard, B., Gao, Z., and Rapoport, N., Am.<br />

J. Drug Deliv., in press.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 435<br />

The aqueous base of the drug formulation, reduced systemic toxicity, potentials for tumor<br />

targeting, and on-demand drug release triggered <strong>by</strong> ultrasound are important additional advantages<br />

of micellar formulations of PTX.<br />

21.4.4.3 Poorly Vascularized Tumors<br />

A different type of drug resistance was observed for subcutaneously grown colon cancer HCT116<br />

tumors. These tumors demonstrated very loose capillary networks and large necrotic areas filled<br />

with ascetic fluid. HCT116 colon cancer tumors manifested strong resistance to the intravenous<br />

drug injections independent on the delivery modality (Figure 21.13, upper set of curves). The<br />

resistance was caused <strong>by</strong> poor vascularization of the tumor volume, resulting in insufficient drug<br />

delivery to tumor cells.<br />

Poor tumor vascularization represents a unique type of tumor resistance to chemotherapy. Upon<br />

intravenous drug injections, avascular tumor regions receive insufficient levels of chemotherapeutic<br />

agents; cancerous cells in these regions remain viable and promote tumor growth. This type of<br />

resistance has not been widely discussed in the literature.<br />

In our laboratory, successful treatment of HCT116 tumors was achieved <strong>by</strong> direct intratumoral<br />

injections of micellar-encapsulated DOX (Figure 21.13, lower set of curves). The encapsulated<br />

drug was effectively retained in the tumor volume, with very low, if any, accumulation in the cells<br />

of other organs. In contrast, free DOX quickly leaked from the tumor tissue, and a short time after<br />

the intratumoral injection of free drug, DOX was already found in the heart cells and those of other<br />

organs, causing associated toxicity.<br />

With the intratumoral injections of micellar encapsulated DOX, effective tumor suppression<br />

was observed even without ultrasound (Figure 21.13). We hypothesize that a high hydrostatic<br />

pressure generated <strong>by</strong> intratumoral injections forced drug-loaded micelles through cell membranes.<br />

For therapy of poorly vascularized tumors, intratumoral injections of micellar-encapsulated<br />

drugs present a viable alternative to inefficient intravenous injections.<br />

The advantages of intratumoral injections were also observed for highly vascularized A2780<br />

tumors. The beneficial effect of ultrasound was revealed in drug biodistribution experiments. For<br />

the intravenous injections of micellar-encapsulated DOX, tumor sonication resulted in a<br />

Tumor volume, mm<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

0 20<br />

40<br />

Days after first treatment<br />

Intravenous<br />

Injections<br />

Intratumaoral<br />

Injections<br />

60 80 100<br />

FIGURE 21.13 Growth curves of HCT116 tumors upon the intravenous (upper group of curves) or intratumoral<br />

(lower group of curves) injections of DOX at a dose rate of 3 mg/kg. Intravenous injections: filled<br />

diamonds—control, open triangles—DOX/PBS, open squares—DOX/mixed micelles/ultrasound; Intratumoral<br />

injections: open circles—DOX/mixed micelles, closed circles—DOX/mixed micelles/ultrasound.<br />

(From Gao, Z. and Rapoport, N., unpublished results.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


436<br />

dramatically increased intracellular drug uptake; for the intratumoral injections of the same formulation,<br />

the main effect of ultrasound was related to a much more uniform drug distribution over the<br />

tumor cell population (Figure 21.14).<br />

Intratumoral chemotherapy <strong>by</strong> nanoparticle encapsulated drugs could be advantageous for<br />

tumors with well-defined primary lesions such as breast, colorectal, prostate, and skin cancers.<br />

These procedures may be performed noninvasively or minimally invasively through intraluminal,<br />

laparoscopic or percutaneous means under the guidance of an appropriate imaging modality such as<br />

MRI, CT, or ultrasound. During the last decade, the interest in intratumoral chemotherapy has<br />

significantly increased. 68<br />

21.4.5 MECHANISMS OF MICELLE/ULTRASOUND BIOEFFECTS<br />

Ultrasound irradiation may play multiple roles in tumor therapy. Ultrasound may (1) induce thermal<br />

effects, (2) enhance extravasation of drug-loaded micelles into the tumor interstitium, (3) enhance<br />

drug diffusion through the tumor interstitium, (4) trigger drug release from micelles in the tumor<br />

volume, and (5) enhance the intracellular uptake of both released and encapsulated drug at the<br />

irradiation site. Though the relative importance of these factors remains to be further explored,<br />

some conclusions can be drawn based on the results presented above.<br />

Drug release from perturbed micelles under the action of ultrasound is undoubtedly an important<br />

factor in the ultrasound-enhanced drug uptake <strong>by</strong> the tumor cells from micellar carriers.<br />

Ultrasound can also enhance the intracellular uptake of intact drug-loaded micelles due to cell<br />

membrane perturbation. Another important component of the ultrasound effect is associated with<br />

the ultrasound-enhanced diffusion of free or micellar-encapsulated drug over tumor tissue.<br />

In summary, drug encapsulation in polymeric micelles combined with local ultrasonic tumor<br />

irradiation results in effective tumor targeting and suppression of growth of drug-sensitive and<br />

multidrug resistant tumors. The effect is associated with the ultrasound-enhanced cytotoxic action<br />

of passively targeted or intratumorally injected micellar-encapsulated chemotherapeutic agents.<br />

(a)<br />

(b)<br />

Counts<br />

Counts<br />

10 0<br />

10 0<br />

1<br />

10 1<br />

10 1<br />

10 2<br />

10 2<br />

Fluorescence, arb.u.<br />

10 3<br />

1 2<br />

FIGURE 21.14 Fluorescence histograms of A2780 tumor cells in (a) unsonicated, and (b) sonicated mouse;<br />

DOX encapsulated in stabilized Pluronic micelles was injected at a dose of at 6 mg/kg (1) intravenously, and<br />

(2) intratumorally; sonicated tumors were irradiated for 30 s <strong>by</strong> 1-MHz ultrasound at 3.4 W/cm 2 power density<br />

and 50% duty cycle 4 h after the drug injection. (From Gao, Z. and Rapoport, N., unpublished results.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

2<br />

Nanotechnology for Cancer Therapy<br />

10 3<br />

10 4<br />

10 4


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 437<br />

This allows on-demand drug delivery and control of systemic, local, and subcellular<br />

drug distribution.<br />

21.5 COMBINING ULTRASOUND IMAGING AND TREATMENT<br />

Ultrasonic tumor treatment requires tumor imaging prior to therapy. Using ultrasound for both<br />

tumor imaging and treatment is especially attractive. Development of ultrasound contrast agents<br />

and concomitant use of harmonic ultrasound imaging has recently made possible imaging of liver, 69<br />

breast, 70 and prostate tumors. 71 Ultrasound-stimulated acoustic emission of microbubbles was used<br />

for color Doppler imaging of liver metastases. 72<br />

Some dual ultrasonic imaging/treatment techniques are already in clinical practice, others are in<br />

the investigational stage. Examples include Sonablate w for prostate cancer treatment (Focus<br />

Surgery, Inc., Indianapolis, Indiana) and a miniature ultrasound-based dual-modality device for<br />

image guided soft tissue ablation (Guided Therapy Systems, Mesa, Arizona).<br />

With dual-modality ultrasound instruments already on the market, developing dual-modality<br />

ultrasound imaging/drug carrier systemswould be very timely.<br />

21.5.1 DUAL-MODALITY CONTRAST AGENTS/DRUG CARRIER SYSTEMS<br />

Ultrasonically enhanced site-specific drug delivery using acoustically activated drug-loaded microbubbles<br />

was suggested <strong>by</strong> Ungar 73 and Ferrara; 74–77 these authors developed echogenic<br />

microbubbles stabilized <strong>by</strong> lipid bilayers. Drug was incorporated in the oil layer located on the<br />

microbubble surface. An effective ultrasound-triggered transfer of a solute from the modified<br />

lipospheres to the tumor cells has been demonstrated. Polymeric ultrasound contrast agents with<br />

targeting potentials have also been explored <strong>by</strong> the Wheatley group. 78–81<br />

Nano/microbubbles oscillate in the ultrasound field and can be cavitated for site-specific localized<br />

delivery of chemotherapeutic agents; they concentrate ultrasound energy and may be used for<br />

enhancing drug delivery from micelles and other nanoparticles. Nano/microemulsions and microbubbles<br />

are efficient reflectors of ultrasound energy, which may be used for contrast-enhanced<br />

ultrasound imaging of tumor neovasculature.<br />

We have recently developed a new class of these agents that combine diagnostic and therapeutic<br />

properties and can be used for ultrasonically enhanced drug delivery and tumor imaging. Our<br />

microbubbles differ in a number of ways from the products on the market or those under development<br />

<strong>by</strong> other groups:<br />

† The microbubbles are produced in situ upon injection of specially designed microemulsions.<br />

† The microbubbles have strong walls produced <strong>by</strong> biodegradable diblock copolymers; the<br />

walls stabilize microbubbles for ultrasound imaging.<br />

† The diblock copolymer not only forms microbubble walls but also forms polymeric<br />

micelles that effectively encapsulate chemotherapeutic agents and act as drug carriers.<br />

† Upon injection, a localized ultrasonic irradiation of the tumor in the presence of microbubbles<br />

provides for effective intracellular drug uptake <strong>by</strong> the tumor cells, which results<br />

in inhibition of tumor growth.<br />

Echogenic nano/microbubbles designed to interact specifically with neovascular epithelium<br />

may allow ultrasound imaging of small metastases. To ensure molecular targeting, the nano/microparticle<br />

surface should be decorated <strong>by</strong> site-specific ligands. In the future, we anticipate conjugating<br />

block copolymer molecules with vectors that would target micelles and nano/microbubbles to the<br />

receptors that are overexpressed on the endothelial lining of tumor neovasculature.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


438<br />

21.6 CONCLUSIONS<br />

In animal models, drug encapsulation within polymeric micelles combined with ultrasonic tumor<br />

irradiation results in effective tumor targeting and inhibition of tumor growth. The effect is based on<br />

ultrasonically triggered localized drug release from micelles and enhanced intracellular drug<br />

uptake. This technique offers prospects for treating multidrug resistant tumors that fail conventional<br />

chemotherapy. Dual-modality drug delivery/ultrasound contrast agent systems used with dualmodality<br />

ultrasound imaging/treatment instruments are expected to allow precisely controlled,<br />

ultrasound-enhanced tumor chemotherapy in a clinical environment.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Kwon, G. S. and Kataoka, K., Block copolymer micelles as long circulating drug vehicles, Drug<br />

Deliv. Rev., 16, 295, 1995.<br />

2. Aliabadi, H. and Lavasanifar, A., Polymeric micelles for drug delivery, Expert Opin. Drug Deliv., 3,<br />

139, <strong>2006</strong>.<br />

3. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent<br />

smancs, Cancer Res., 46, 6387, 1986.<br />

4. Zhang, X., Burt, H. M., and Von Hoff, D., An investigation of the antitumor activity and biodistribution<br />

of polymeric micellar paclitaxel, Cancer Chemother. Pharmacol., 40, 81, 1997.<br />

5. Zhang, X., Jackson, J. K., and Burt, H. M., Development of amphiphilic diblock copolymers as<br />

micellear carriers of taxol, Int. J. Pharm., 132, 195, 1996.<br />

6. Zhang, X., Burt, H. M., and Mangold, G., Anti-tumor efficacy and biodistribution of intravenous<br />

polymeric micellar paclitaxel, Anticancer Drug, 8, 696, 1997.<br />

7. Adams, M. L., Andes, D. R., and Kwon, G. S., Amphotericin B encapsulated in micelles based on<br />

poly(ethylene oxide)-b-poly(L-amino acid) derivatives exerts reduced in vitro hemolysis but maintains<br />

potent in vivo antifungal activity, Biomacromolecules, 4, 750, 2003.<br />

8. Adams, M. L. and Kwon, G. S., Relative aggregation state and hemolytic activity of amphotericin B<br />

encapsulated <strong>by</strong> poly(ethylene oxide)-b-poly(N-hexyl-L-aspartamide)-acyl conjugate micelles:<br />

Effects of acyl chain length, J. Control. Release, 87, 23, 2003.<br />

9. Leung, S., Jackson, J., Miyake, H., Burth, H., and Gleave, M. E., Polymeric micellar paclitaxel<br />

phosphorylates Bcl-2 and induces apopototic regression of androgen-independent LNCaP prostate<br />

tumors, Prostate, 44, 156, 2000.<br />

10. Shin, I. G. et al., Methoxy poly(ethylene glycol)/epsilon-caprolactone amphiphilic block copolymeric<br />

micelle containing indomethacin. I. Preparation and characterization, J. Control. Release, 51, 1, 1998.<br />

11. Kim, S. C. et al., In vivo evaluation of polymeric micellar paclitaxel formulation: Toxicity and<br />

efficacy, J. Control. Release, 72, 191, 2001.<br />

12. Benahmed, A., Ranger, M., and Leroux, J. C., Novel polymeric micelles based on the amphiphilic<br />

diblock copolymer poly(N-vinyl-2-pyrrolidone)-b-poly(D,L-lactide), Pharm. Res., 18, 323, 2001.<br />

13. Yu, B. G. et al., Polymeric micelles for drug delivery: Solubilization and haemolytic activity of<br />

amphotericin B, J. Control. Release, 53, 131, 1998.<br />

14. Lavasanifar, A., Samuel, J., and Kwon, G. S., Micelles of poly(ethylene oxide)-b-poly(N-alkyl stearate<br />

L-aspartamide): Synthetic analogues of lipoproteins for drug delivery, J. Biomed. Mater. Res., 52,<br />

831, 2000.<br />

15. Piskin, E., Kaitian, X., Denkbas, E. B., and Kucukyavuz, Z., Novel PDLLA/PEG copolymer micelles<br />

as drug carriers, J. Biomater. Sci. Polym. Ed., 7, 359, 1995.<br />

16. <strong>Francis</strong>, M. F., Piredda, M., and Winnik, F. M., Solubilization of poorly water soluble drugs in<br />

micelles of hydrophobically modified hydroxypropylcellulose copolymers, J. Control. Release, 93,<br />

59, 2003.<br />

17. Jette, K. K. et al., Preparation and drug loading of poly(ethylene glycol)-b-poly(epsilon-caprolactone)<br />

micelles through the evaporation of a cosolvent azeotrope, Pharm. Res., 21, 1184, 2004.<br />

18. Nakanishi, T. et al., Development of the polymer micelle carrier system for doxorubicin, J. Control.<br />

Release, 74, 295, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 439<br />

19. Rapoport, N., Stabilization and acoustic activation of pluronic micelles for tumor-targeted drug<br />

delivery, Colloids Surf. B Biointerfaces, 3, 93, 1999.<br />

20. Shuai, X. et al., Core-cross-linked polymeric micelles as paclitaxel carriers, Bioconj. Chem., 15, 441,<br />

2004.<br />

21. Lee, J., Cho, E. C., and Cho, K., Incorporation and release behavior of hydrophobic drug in functionalized<br />

poly(D,L-lactide)-b-poly(ethylene oxide) micelles, J. Control. Release, 94, 323, 2004.<br />

22. Matsumura, Y. et al., Phase I clinical trial and pharmacokinetic evaluation of NK911, a micelleencapsulated<br />

doxorubicin, Br. J. Cancer, 91, 1775, 2004.<br />

23. Alakhov, V. et al., Hypersensitization of multidrug resistant human ovarian carcinoma cells <strong>by</strong><br />

pluronic P85 block copolymer, Bioconjug. Chem., 7, 209, 1996.<br />

24. Batrakova, E. V. et al., Anthracycline antibiotics non-covalently incorporated into the block copolymer<br />

micelles: In vivo evaluation of anti-cancer activity, Br. J. Cancer, 74, 1545, 1996.<br />

25. Venne, A. et al., Hypersensitizing effect of pluronic L61 on cytotoxic activity, transport, and subcellular<br />

distribution of doxorubicin in multiple drug-resistant cells, Cancer Res., 56, 3626, 1996.<br />

26. Batrakova, E. V. et al., Effects of pluronic block copolymers on drug absorption in Caco-2 cell<br />

monolayers, Pharm. Res., 850, 1998.<br />

27. Batrakova, E. V. et al., Effects of pluronic P85 unimers and micelles on drug permeability in polarized<br />

BBMEC and Caco-2 cells, Pharm. Res., 15, 1525, 1998.<br />

28. Miller, D. W., Batrakova, E. V., and Kabanov, A. V., Inhibition of multidrug resistance-associated<br />

protein (MRP) functional activity with pluronic block copolymers, Pharm. Res., 16, 396, 1999.<br />

29. Batrakova, E. et al., Fundamental relationships between the composition of pluronic block copolymers<br />

and their hypersensitization effect in MDR cancer cells, Pharm. Res., 16, 1373, 1999.<br />

30. Alakhov, V. et al., Block copolymer-based formulation of doxorubicin. From cell screen to clinical<br />

trials, Colloids Surf. B Biointerfaces, 16, 113, 1999.<br />

31. Batrakova, E. V. et al., Optimal structure requirements for pluronic block copolymers in modifying<br />

P-glycoprotein drug efflux transporter activity in bovine brain microvessel endothelial cells,<br />

J. Pharmacol. Exp. Ther., 304, 845, 2003.<br />

32. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., An essential relationship between ATP<br />

depletion and chemosensitizing activity of Pluronic block copolymers, J. Control. Release, 91, 75,<br />

2003.<br />

33. Venne, A. et al., Hypersensitizing effect of pluronic L61 on cytotoxic activity, transport, and subcellular<br />

distribution of doxorubicin in multiple drug-resistant cells, Cancer Res., 56, 3626, 1996.<br />

34. Rapoport, N. et al., Intracellular uptake and trafficking of Pluronic micelles in drug-sensitive and<br />

MDR cells: Effect on the intracellular drug localization, J. Pharm. Sci., 91, 157, 2002.<br />

35. Danson, S. et al., Phase I dose escalation and pharmacokinetic study of pluronic polymer-bound<br />

doxorubicin (SP1049C) in patients with advanced cancer, Br. J. Cancer, 90, 2085, 2004.<br />

36. Danson, S. et al., Phase I dose escalation and pharmacokinetic study of pluronic polymer-bound<br />

doxorubicin (SP1049C) in patients with advanced cancer, Br. J. Cancer, 90, 2085, 2004.<br />

37. Valle, J. W. et al., A phase II, window study of SP1049C as first-line therapy in inoperable metastatic<br />

adenocarcinoma of the esophagus, J. Clin. Oncol., 22, 4195, 2004.<br />

38. Gelderblom, H. et al., Cremophor EL: The drawbacks and advantages of vehicle selection for drug<br />

formulation, Eur. J. Cancer, 37, 1590, 2001.<br />

39. Kim, S. C. et al., In vivo evaluation of polymeric micellar paclitaxel formulation: Toxicity and<br />

efficacy, J. Control. Release, 72, 191, 2001.<br />

40. Kim, T. Y. et al., Phase I and pharmacokinetic study of Genexol-PM, a cremophor-free, polymeric<br />

micelle-formulated paclitaxel, in patients with advanced malignancies, Clin. Cancer Res., 10, 3708,<br />

2004.<br />

41. Fournier, E. et al., A novel one-step drug-loading procedure for water-soluble amphiphilic nanocarriers,<br />

Pharm. Res., 21, 962, 2004.<br />

42. Le Garrec, D. et al., Poly(N-vinylpyrrolidone)-b-poly(D,L-lactide) as a new polymeric solubilizer for<br />

hydrophobic anticancer drugs: In vitro and in vivo evaluation, J. Control. Release, 99, 83, 2004.<br />

43. Hamaguchi, T. et al., NK105, a paclitaxel-incorporating micellar nanoparticle formulation, can extend<br />

in vivo antitumour activity and reduce the neurotoxicity of paclitaxel, Br. J. Cancer, 92, 1240, 2005.<br />

44. Gao, Z. G., Fain, H. D., and Rapoport, N., Controlled and targeted tumor chemotherapy <strong>by</strong> micellarencapsulated<br />

drug and ultrasound, J. Control. Release, 102, 203, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


440<br />

Nanotechnology for Cancer Therapy<br />

45. Marin, A., Muniruzzaman, M., and Rapoport, N., Mechanism of the ultrasound activation of micellar<br />

drug delivery, J. Control. Release, 75, 69, 2001.<br />

46. Husseini, G. A. et al., Factors affecting acoustically triggered release of drugs from polymeric<br />

micelles, J. Control. Release, 69, 43, 2000.<br />

47. Gao, Z., Fain, H. D., and Rapoport, N., Ultrasound-enhanced tumor targeting of polymeric micellar<br />

drug carriers, Mol. Pharm., 1, 317, 2004.<br />

48. Rapoport, N., Tumor targeting <strong>by</strong> polymeric assemblies and ultrasound activation, In Kentus Books<br />

MML Series, Arshadi, R. and Kono, K., Eds., Vol. 8, Kentus Books, London, pp. 305–362, 2005.<br />

49. Rapoport, N., Combined cancer therapy <strong>by</strong> micellar-encapsulated drug and ultrasound, Int. J. Pharm.,<br />

277, 155, 2004.<br />

50. Rapoport, N., Marin, A., and Christensen, D. A., Ultrasound-activated drug delivery, Drug Deliv. Syst.<br />

Sci., 2, 37, 2002.<br />

51. Rapoport, N., Factors affecting ultrasound interactions with polymeric micelles and viable cells, In<br />

Carrier-Based Drug Delivery, ACS Symposium Series, Swenson, S., Ed., Vol. 879, Chapter 12,<br />

American Chemical Society, Washington, DC, 2004.<br />

52. Rapoport, N., Marin, A., Muniruzzaman, M., and Christensen, D., Controlled drug delivery to drugsencitive<br />

and multidrug resistant cells: Effects of pluronic micelles and ultrasound, In Advances in<br />

Controlled Drug Delivery, ACS Symposium Book Series, Dinh, S. M. and Liu, P., Eds., American<br />

Chemical Society, Washington, DC, pp. 85–101, 2003.<br />

53. Rapoport, N. et al., Drug delivery in polymeric micelles: From in vitro to in vivo, J. Control. Release,<br />

91, 85, 2003.<br />

54. Rapoport, N. Y. et al., Ultrasound-triggered drug targeting of tumors in vitro and in vivo, Ultrasonics,<br />

42, 943, 2004.<br />

55. Husseini, G. A. et al., Kinetics of ultrasonic release of Doxorubicin from Pluronic P-105 micelles,<br />

Colloids Surf. B Biointerfaces, 24, 253, 2002.<br />

56. Marin, A. et al., Drug delivery in pluronic micelles: Effect of high-frequency ultrasound on drug<br />

release from micelles and intracellular uptake, J. Control. Release, 84, 39, 2002.<br />

57. Deng, C. X. et al., Ultrasound-induced cell membrane porosity, Ultrasound Med. Biol., 30, 519, 2004.<br />

58. Honda, H. et al., Role of intracellular calcium ions and reactive oxygen species in apoptosis induced<br />

<strong>by</strong> ultrasound, Ultrasound Med. Biol., 30, 683, 2004.<br />

59. Pan, H. et al., Study of sonoporation dynamics affected <strong>by</strong> ultrasound duty cycle, Ultrasound Med.<br />

Biol., 31, 849, 2005.<br />

60. Miller, D. L. and Quddus, J., Sonoporation of monolayer cells <strong>by</strong> diagnostic ultrasound activation of<br />

contrast-agent gas bodies, Ultrasound Med. Biol., 26, 661, 2000.<br />

61. Kamaev, P. P. and Rapoport, N. Y., Effect of anticancer drug on the cell sensitivity to ultrasound in<br />

vitro and in vivo, AIP Conf. Proc., 829, 543, <strong>2006</strong>.<br />

62. Li, P., Armstrong, W. F., and Miller, D. L., Impact of myocardial contrast echocardiography on<br />

vascular permeability: Comparison of three different contrast agents, Ultrasound Med. Biol., 30,<br />

83, 2004.<br />

63. Li, P. et al., Impact of myocardial contrast echocardiography on vascular permeability: An in vivo<br />

dose response study of delivery mode, pressure amplitude and contrast dose, Ultrasound Med. Biol.,<br />

29, 1341, 2003.<br />

64. Miller, D. L. et al., Histological characterization of microlesions induced <strong>by</strong> myocardial contrast<br />

echocardiography, Echocardiography, 22, 25, 2005.<br />

65. Wible, J. H. et al., Microbubbles induce renal hemorrhage when exposed to diagnostic ultrasound in<br />

anesthetized rats, Ultrasound Med. Biol., 28, 1535, 2002.<br />

66. Bednarski, M. D. et al., In vivo target-specific delivery of macromolecular agents with MR-guided<br />

focused ultrasound, Radiology, 204, 263, 1997.<br />

67. Kataoka, K. et al., Doxorubicin-loaded poly(ethylene glycol)-poly(beta-benzyl-L-aspartate)<br />

copolymer micelles: Their pharmaceutical characteristics and biological significance, J. Control.<br />

Release, 64, 143, 2000.<br />

68. Goldberg, E. P. et al., Intratumoral cancer chemotherapy and immunotherapy: Opportunities for<br />

non-systemic preoperative drug delivery, J. Pharm. Pharmacol., 54, 159, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Combined Cancer Therapy <strong>by</strong> Micellar-Encapsulated Drugs and Ultrasound 441<br />

69. Wilson, S. R. et al., Harmonic hepatic US with microbubble contrast agent: Initial experience showing<br />

improved characterization of hemangioma, hepatocellular carcinoma, and metastasis, Radiology, 215,<br />

153, 2000.<br />

70. Kedar, R. P. et al., Microbubble contrast agent for color Doppler US: Effect on breast masses. Work in<br />

progress, Radiology, 198, 679, 1996.<br />

71. Halpern, E. J., Rosenberg, M., and Gomella, L. G., Prostate cancer: Contrast-enhanced ultrasound for<br />

detection, Radiology, 219, 219, 2001.<br />

72. Bauer, A. et al., Ultrasonic imaging of organ perfusion with SH U 563A, Acad. Radiol., 9(Suppl 1),<br />

S46, 2002.<br />

73. Unger, E. C. et al., Therapeutic applications of lipid-coated microbubbles, Adv. Drug Deliv. Rev., 56,<br />

1291, 2004.<br />

74. Shortencarier, M. J. et al., A method for radiation-force localized drug delivery using gas-filled<br />

lipospheres, IEEE Trans. Ultrason. Ferroelectr. Freq. Control, 51, 822, 2004.<br />

75. Bloch, S. H., Dayton, P. A., and Ferrara, K. W., Targeted imaging using ultrasound contrast agents.<br />

Progess and opportunities for clinical and research applications, IEEE Eng. Med. Biol. Mag., 23, 18,<br />

2004.<br />

76. Bloch, S. H. et al., The effect of size on the acoustic response of polymer-shelled contrast agents,<br />

Ultrasound Med. Biol., 31, 439, 2005.<br />

77. Dayton, P. A. and Ferrara, K. W., Targeted imaging using ultrasound, J. Magn. Reson. Imaging, 16,<br />

362, 2002.<br />

78. Lathia, J. D., Leodore, L., and Wheatley, M. A., Polymeric contrast agent with targeting potential,<br />

Ultrasonics, 42, 763, 2004.<br />

79. El-Sherif, D. M. et al., Ultrasound degradation of novel polymer contrast agents, J. Biomed. Mater.<br />

Res. A, 68, 71, 2004.<br />

80. El-Sherif, D. M. and Wheatley, M. A., Development of a novel method for synthesis of a polymeric<br />

ultrasound contrast agent, J. Biomed. Mater. Res. A, 66, 347, 2003.<br />

81. Basude, R. and Wheatley, M. A., Generation of ultraharmonics in surfactant based ultrasound contrast<br />

agents: Use and advantages, Ultrasonics, 39, 437, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


22<br />

CONTENTS<br />

Polymeric Micelles Targeting<br />

Tumor pH<br />

Eun Seong Lee and You Han Bae<br />

22.1 Introduction ....................................................................................................................... 443<br />

22.2 Acidic Tumor pH e ............................................................................................................. 444<br />

22.3 Multidrug Resistance and Cancer ..................................................................................... 444<br />

22.4 PolyHis-PEG-Based pH-Sensitive Micelles: Preparation and Characteristics ................ 445<br />

22.4.1 Synthesis .............................................................................................................. 445<br />

22.4.2 PolyHis/PEG Copolymer Micelles ..................................................................... 446<br />

22.4.3 Mixed Micelles of PolyHis/PEG and PLLA/PEG.............................................. 448<br />

22.5 Tumor pHe Targeting ........................................................................................................ 451<br />

22.6 Early Endosomal pH Targeting for Reversal of MDR .................................................... 455<br />

22.7 Conclusions ....................................................................................................................... 460<br />

References..................................................................................................................................... 461<br />

22.1 INTRODUCTION<br />

The dose–response curves for cell cycle-phase nonspecific anti-cancer drugs, such as doxorubicin,<br />

have been constructed from in vitro experimental results. The survival fraction of both wild-type<br />

sensitive and resistant cancer cells decreases in a sigmoidal pattern as a function of log of dose size<br />

(the survival fraction of the cells remains the same as the untreated control at low concentrations,<br />

starts to sharply decrease above a specific concentration and reaches a plateau at high concentrations).<br />

The inflection point of the curves is drug specific and is shifted to a lower concentration<br />

with an increasing duration of drug exposure. The inflection point for the resistant cells occurs at a<br />

higher concentration region and shows a more gradual decrease in viability with increasing drug<br />

concentration. 1,2 Overall, the experimental results and theoretical fitting of these results give us a<br />

simple implication that high dose exposure for a longer duration results in effective destruction.<br />

Two classical nanosized anti-cancer drug carriers are liposomes and polymeric micelles. Each<br />

carrier system shows advantages as well as disadvantages over the other. For instance, liposomal<br />

carriers show continuous leakage of an entrapped drug with instability during storage and circulation.<br />

Polymeric micelles demonstrate much improved stability, but they may still dissociate upon<br />

dilution after injection and interact with blood components. 3,4<br />

Considering the dose–response relationship, it was suggested that future improvement in carrier<br />

design, providing a triggerable mechanism for drug release upon reaching the tumor sites, could be<br />

the most efficacious delivery strategy. 5 Various approaches for externally stimulated triggered<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

443


444<br />

release have been attempted, including hyperthermic conditions 6 used for thermosensitive liposomes<br />

or polymeric micelles, and sonication 7 used to trigger release from Pluronic w micelles.<br />

In this chapter, polymeric micelles for the triggered release either <strong>by</strong> slightly acidic extracellular<br />

pH (pH e) or <strong>by</strong> early endosomal pH, are introduced. Triggered release <strong>by</strong> tumor pH e may allow<br />

localized high-dose therapy, which is effective for sensitive cells. Active intracellular translocation<br />

of the micellar carriers via specific interactions, combined with triggered release in early endosomes<br />

and endosomolytic activity of destabilized micelle components, has been proven to be<br />

effective for treating multidrug-resistant tumors.<br />

22.2 ACIDIC TUMOR pHe<br />

The pHe of normal tissues and blood is around pH 7.4, and the intracellular pH (pHi) of normal cells<br />

is around pH 7.2. However, in most tumor cells the pH gradient is reversed (pHiOpHe). Particularly,<br />

pHe is lower than in normal tissues. 8–11 Although there is a distribution of in vivo pHe measurements of human patients having various solid tumors (adenocarcinoma, squamous cell<br />

carcinoma, soft tissue sarcoma, and malignant melanoma) in readily accessible areas (limbs,<br />

neck, or chest wall) using needle type microelectrodes, the mean pH value is reported to be 7.0<br />

with a full range of 5.7–7.8, 11 the variation being dependent on tumor histology and tumor volume.<br />

Recent measurements of pHe <strong>by</strong> noninvasive technologies, such as 19 F, 31 P, or 1 Hprobesin<br />

magnetic resonance spectroscopy in human-tumor xenografts in animals, further verified consistently<br />

low pHe. 12,13 All measurements of pHe of human and animal solid-tumors <strong>by</strong> either invasive<br />

or noninvasive method showed that more than 80% of all measured values fall below pH 7.2. This<br />

tumor pH can be further lowered <strong>by</strong> hyperglycemic conditions. 14<br />

The high rate of glycolysis in tumor cells, under either aerobic or anaerobic conditions, has<br />

been thought to be a major cause of low pHe. The tumor cells synthesize ATP both <strong>by</strong> mitochondrial<br />

oxidative phosphorylation and glycolysis. Glycolysis produces two moles of lactic acid (pKaZ3.9)<br />

and two moles of ATP upon consumption of one mole of glucose. The hydrolysis of generated ATP<br />

produces protons. Despite the high production rate of protons in tumor cells, their cytosolic pH,<br />

particularly of resistant cells, is maintained to be alkaline, creating a favorable condition for<br />

glycolysis. The mechanisms responsible for exporting protons out of the cells remain unknown.<br />

This, along with the help of inadequate blood supply, poor lymphatic drainage, and high interstitial<br />

pressure in tumor tissues, leads to low pHe. 8–10 However, it is of interest to note that in one study,<br />

glycolysis-deficient variant cells (lacking lactate dehydrogenase) produced negligible quantities of<br />

lactic acid but still presented an acidic environment. 15 This finding has indicated the acidity may be<br />

a phenotype of tumor cells rather than a consequence of cell metabolic events. The acidic environment<br />

may benefit the cancer cell because it promotes invasiveness (metastasis) <strong>by</strong> destroying the<br />

extracellular matrix of surrounding normal tissues.<br />

A few attempts for pH-triggered release using pH-sensitive liposomes have been described in<br />

the literature, 16,17 but a practical system that triggers the release at tumor pHe is far off because of<br />

technical difficulties in endowing the desired pH sensitivity to the liposomes.<br />

22.3 MULTIDRUG RESISTANCE AND CANCER<br />

Nanotechnology for Cancer Therapy<br />

The resistance of cancer cells to multiple structurally unrelated chemotherapeutic drugs has been<br />

termed multidrug resistance (MDR). Clinical resistance of some cancers to cytotoxic drugs has<br />

been observed since the beginning of chemotherapy. The degree of resistance varies widely<br />

depending on the type of cancer. In past years, various mechanisms for MDR have been proposed,<br />

including the expression of cell surface multidrug efflux pumps (P-glycoprotein (Pgp) 18 and<br />

multidrug resistance-associated proteins (MRP1) 19 ) altered glutathione metabolism, 20 reduced<br />

activity of topoisomerase II, 21 and various changes in cellular proteins 22–24 and mechanisms. 25<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 445<br />

Strong evidence shows that the expression of (Pgp) is associated with drug-resistance in cancer.<br />

Many attempts to circumvent Pgp-based MDR in cancer chemotherapy have utilized Pgp blocking<br />

agents (MDR modulators or reverters), 26–30 that can be co-administered with the anti-cancer drug.<br />

This approach is based on the premise that inhibiting Pgp function will result in increased accumulation<br />

of most anti-cancer drugs in the tumor cells and restore anti-tumor activity. However,<br />

co-administration of MDR modulators with anti-cancer drugs has often resulted in exacerbated<br />

toxicity of the anti-cancer drugs and very little improvement in chemosensitization of MDR cancers<br />

because of the blockade of Pgp excretory functions in healthy tissues, such as liver and kidney,<br />

which markedly reduces anti-cancer drug clearance.<br />

Independently, acidic intracellular organelles also participate in resistance to chemotherapeutic<br />

drugs. Parental drug sensitive cells are characterized to have rather acidic and diffused pH profile<br />

inside the cells, while multidrug resistant cells develop more acidic organelles (recycling endosome,<br />

lysosome, and trans-Golgi network) compared to cytosolic and nucleoplasmic pH. 31–33<br />

Because most anti-cancer drugs are in an ionizable form, the pH of extracellular matrix and<br />

intracellular compartments is a critical factor in determining drug partitioning and distribution. For<br />

instance, weakly-basic drugs accumulate more in an acidic phase, where the ionized form predominates,<br />

because an ionized drug is less permeable through cellular or subcellular membranes. This<br />

is more pronounced in MDR cells.<br />

It is claimed that the drug sequestration followed <strong>by</strong> transport to the plasma membrane and<br />

extrusion into the external medium is another major mechanism for MDR for weakly-acidic<br />

drugs. 31,34 It is interesting to note that transfection of P-gp into certain sensitive cells is often<br />

coupled with increased cytosolic pH. However, this is not the case for MCF-7 cell line. The<br />

transfection did not influence the internal pH but induced 20-fold increased resistance. When the<br />

same cell line was selected from an escalated doxorubicin treatment schedule, the cells presented<br />

the overexpression of P-gp together with typical MDR pH profile and increased the resistance 980fold.<br />

These observations indicate that the activity of the efflux pump, combined with drug sequestration<br />

and exocytosis mechanism, seem to be highly influential factors in MDR. 35 Thus, this<br />

subcellular pH effect must be taken into account in the carrier-design to avoid the sequestration<br />

effect and to improve drug availability to the cytosol and nucleus. 34 Few investigators have<br />

addressed the drug sequestration issue in drug-carrier design.<br />

This chapter summarizes our attempt for the triggered release of doxorubicin, either <strong>by</strong> tumor<br />

pHe or <strong>by</strong> early endosomal pH, from pH-sensitive polymeric micelles which are based on poly(Lhistidine)<br />

core.<br />

22.4 POLYHIS-PEG-BASED pH-SENSITIVE MICELLES: PREPARATION AND<br />

CHARACTERISTICS<br />

22.4.1 SYNTHESIS<br />

L-Histidine (His) was derivatized <strong>by</strong> introducing carbobenzoxy (CBZ) group to a amino group, and<br />

amino group in imidazole ring of N a -CBZ-L-histidine was protected with dinitrophenyl (DNP)<br />

group. N a -CBZ-N im -DNP-L-His was then transformed to the N-carboxy anhydride (NCA) form<br />

<strong>by</strong> thionyl chloride. The ring-opening polymerization of N a -CBZ-N im -DNP-L-His NCA$HCl in<br />

dimethylformamide (DMF) was performed using different molar ratios of the monomer to initiator<br />

(isopropylamine or n-hexylamine) (M/I ratio). 36<br />

Monocarboxy-PEG was used to prepare an activated N-hydroxysuccinyl (NHS)–poly(ethylene<br />

glycol) (PEG). The coupling reaction between poly(N im -DNP-L-histidine) and NHS–PEG (1:1<br />

functional group ratio) was carried out in tetrahydrofuran (THF) for two days at room temperature.<br />

After the reaction, the block copolymer was precipitated in a mixed solvent (THF/n-hexane) and<br />

filtered. 2-Mercaptoethanol was added to the poly(N im -DNP-L-histidine)-PEG diblock copolymer<br />

in DMF to deprotect the imidazole group (Scheme 22.1). Deprotection was complete within one<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


446<br />

day at room temperature. The polymer solution was added to diethyl ether (800 mL) at 08C to<br />

precipitate polymer, while excess 2-mercaptoethanol and DNP-mercaptoethanol remained soluble.<br />

The polymer was filtered, washed with diethyl ether, and dried in vacuo for two days. For further<br />

purification, the polymer was dissolved in a minimum volume of 3 N HCl and stored in a 08C for<br />

one day. After the precipitation of DNP was complete, the solution was filtered through a 0.45-mm<br />

filter and the product was precipitated <strong>by</strong> adding an excess amount of acetone to the solution, and<br />

then dried in vacuo for two days. Dialysis (Spectra/Por; molecular weight cut off (MWCO) 5000)<br />

was used to remove uncoupled polymers. Before subsequent experiments, the HCl salt of polyhistidine<br />

(1 mmol) in dimethylsulfoxide (DMSO) was converted to the free base <strong>by</strong> treatment<br />

with TEA (2 mmol) for 3 h at room temperature; the free base was recrystallized from a solution<br />

mixture of ethanol and DMSO (10:1).<br />

It has been documented in literature that histidine residues in proteins significantly contribute to<br />

the protein buffering capacity 37 at physiological pH. The acid-base titration profiles of the polyHis<br />

homopolymer and polyHis/PEG block copolymers are presented in Figure 22.1. All of the polymer<br />

solutions exhibited a buffering pH region of pH 4–9. The titration curve confirmed that the polymers<br />

with a higher molecular-weight polyHis block had a higher buffering capacity (Figure 22.1a)<br />

in the physiological pH range of pH 5.5–8.0 due to the higher concentrations of imidazole group.<br />

PolyHis (M nw5000)/PEG (M nw2000) (PolyHis5K/PEG2K) and PolyHis (M nw3000)/PEG<br />

(M nw2000) (polyHis3K/PEG2K) had inflexion points around pH 7.0 (pK b). PolyHis5K and Poly-<br />

His3K showed pK b values around pH 6.5. The pK b shift of the block copolymer compared to the<br />

polyHis homopolymer may be due to increased hydration <strong>by</strong> PEG. 38<br />

22.4.2 POLYHIS/PEG COPOLYMER MICELLES<br />

The deprotonated polyHis at high pH is hydrophobic, whereas PEG is soluble in water at all pHs.<br />

This amphiphilicity is responsible for the formation of polymeric micelles. Lowering the solution<br />

pH below the pKb can affect the micellar structure because protonation decreases the hydrophobicity<br />

of polyhistidine. 36<br />

The polyHis/PEG block copolymer micelles were prepared <strong>by</strong> the dialysis of the copolymer<br />

solution in DMSO against a pH 8.0 medium. This was done in the presence of pyrene as<br />

H2N CH COOH<br />

CH2 carbobenzyl<br />

chloride<br />

N NH<br />

O<br />

O C C O<br />

HN<br />

CH2 N N<br />

initiator<br />

NO 2<br />

O<br />

H<br />

CH2 O C N CH COOH<br />

CH2 N NH<br />

N α -CBZ-L-histidine<br />

O<br />

H<br />

N CH C initiator residue (R)<br />

CH2 n<br />

N N<br />

NO 2<br />

NO2 NO2 N α -CBZ-N im -DNP-L-histidine NCA poly(N im -DNP-L-histidine)<br />

deblocking<br />

O<br />

PEG C<br />

H<br />

H<br />

O<br />

N CH C R<br />

CH2 n<br />

N NH<br />

PEG-b-poly(L-histidine)<br />

2,4-dinitrofluoro<br />

benzene<br />

carboxy group<br />

activated PEG<br />

Nanotechnology for Cancer Therapy<br />

O<br />

H<br />

CH2 O C N CH COOH<br />

CH2 N N<br />

NO 2<br />

NO2 N α -CBZ-N im -DNP-L-histidine<br />

O<br />

PEG C<br />

N N<br />

thionyl<br />

chloride<br />

O<br />

H<br />

N CH C R<br />

CH2 n<br />

NO 2<br />

NO 2<br />

O<br />

O C C O<br />

HN<br />

CH2 N N<br />

NO 2<br />

NO2 N im -DNP-L-histidine NCA<br />

SCHEME 22.1 Overall scheme for the syntheses of protected L-histidine monomer, protected poly(L-histidine)<br />

<strong>by</strong> ring-opening polymerization of NCA, coupling with PEG, and imidazole ring deprotection. (From<br />

Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., J. Control Release, 90, 363, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 447<br />

pH<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 200 400 600 800<br />

0<br />

0 200 400 600<br />

(a) 1 N HCL (μL)<br />

(b)<br />

1 N HCL (μL)<br />

FIGURE 22.1 The pH profile of (a) polyHis5K/PEG2K (C), polyHis5K (&), NaCl (;) and (b) polyHis3K/<br />

PEG2K (C), polyHis3K (&) <strong>by</strong> acid-base titration. (From Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H.,<br />

J. Control Release, 90, 363, 2003. With permission.)<br />

a fluorescent probe so that the micelle formation could be monitored <strong>by</strong> fluorometry. Pyrene<br />

strongly fluoresces in a nonpolar environment, while in a polar environment it shows weak fluorescence<br />

intensity. The change of total emission intensity vs. polymer concentration indicates the<br />

formation of micelles or the change from micelles to unimers (the dissociatation of polymers from<br />

the micelles). 39 PolyHis3K/PEG2K micelles exhibited instability over time, as evidenced <strong>by</strong><br />

increasing light transmittance, while the stability of polyHis5K/PEG2K micelles was maintained<br />

for two days (data not shown). The instability of polyHis3K/PEG2K was probably due to short<br />

polyHis block length and indicates that there may be a critical polyHis length for stability<br />

somewhere between Mn 3000 and 5000. Therefore, polyHis5K/PEG2K was used for<br />

further investigation.<br />

The effect of the dialysis pH on critical micelle concentration (CMC) was examined and the<br />

results are presented in Figure 22.2. At a dialysis pH between 8.0 and 7.4, the CMC of polyHis5K/<br />

PEG2K micelle increased slightly with decreasing pH. However, the CMC was significantly<br />

elevated below pH 7.2. It is evident that the protonation of the imidazole group in the copolymer<br />

at lower pH causes a reduction in hydrophobicity, leading to an increase in CMC.<br />

CMC (μg/mL)<br />

150<br />

100<br />

50<br />

0<br />

5.0 5.5 6.0<br />

pH<br />

pH<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

6.5 7.0 7.5 8.0<br />

FIGURE 22.2 The pH effect on the CMC of polyHis5K/PEG2K polymeric micelles that were prepared at<br />

various pHs (pH 8.0–5.0). (From Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., J. Control Release, 90, 363,<br />

2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

800


448<br />

Intensity (%)<br />

(a)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1<br />

10 100<br />

Particle size (nm)<br />

1000<br />

In addition, below pH 5.0 (typically pH 4.8), the CMC of polyHis5K/PEG2K micelle could not<br />

be detected. Taken together, at pH 8.0 the less protonated polyHis constitutes the hydrophobic core<br />

in the micellar structure, but in the range of pH 5.0–7.4 the polymer produced less-stable micelles.<br />

The size of the micelles formed from polyHis5K/PEG2K measured <strong>by</strong> a zetasizer was around<br />

110 nm, based on the intensity-average diameter with a unimodal distribution (Figure 22.3a), and<br />

the ratio of weight-average particle size to number-average particle size was 1.2. The relatively<br />

large size of polyHis5K/PEG2K micelles may be the result of dialysis conditions, which probably<br />

produced a secondary micelle structure that is a cluster of individual single micelles. 40 However,<br />

the polyHis5K/PEG2K micellar shape was regular and spherical, as visualized in the atomic force<br />

microscopy photograph (Figure 22.3b). The morphology of polymeric micelles was further evident<br />

from dynamic light scattering (DLS) measurements.<br />

A noticeable increase in light-scattering intensity was observed after dialyzing the block copolymer<br />

solutions at pH 8.0, while the unimer solution in DMSO was practically transparent. The<br />

pH-dependent micelle stability was confirmed again <strong>by</strong> the measurement of transmittance of the<br />

micelle solution. Figure 22.4a shows the transmittance change of the polymeric micelle solution as<br />

a function of pH. As the pH of the micelle solution decreased from the dialysis conditions (pH 8,<br />

ionic strength 1.0), the transmittance increased. In particular, this increase was prominent from pH<br />

7.4 to 6.0, reaching a plateau around pH 6.4. This result implies that polyHis5K/PEG2K micelles<br />

start dissociation starting at pH 7.4 and show a sharp transition between pH 7.4 and 7.0 followed <strong>by</strong> a<br />

gradual transition between pH 7.0 and 6.0.<br />

Figure 22.4b shows the changes in the particle size distribution of the micelles as a function of<br />

pH. The instability of the micelles on reduction of pH resulted in release of polyHis5K/PEG2K<br />

unimers or unimicelles isolated from secondary micelles, resulting in a bimodal size distribution<br />

with existing secondary micelles. The intensity of the peak at size 2–30 nm gradually increased<br />

with micelle dissociation at a lower pH. Below pH 7.4, an abrupt increase in the intensity of smallsized<br />

particles was apparent, which is consistent with the transmittance change of the micelle<br />

solution at this pH (Figure 22.4a).<br />

22.4.3 MIXED MICELLES OF POLYHIS/PEG AND PLLA/PEG<br />

The PolyHis/PEG micelle was relatively unstable and began to gradually disintegrate at pH 7.4.<br />

The incorporation of a nonionizable block copolymer (Poly(L-lactic acide) (PLLA)/PEG) into the<br />

Å<br />

20.0<br />

10.0<br />

0.0<br />

(b)<br />

Nanotechnology for Cancer Therapy<br />

0 0.2 0.4 µm<br />

FIGURE 22.3 The particle size distribution of polyHis5K/PEG2K polymeric micelles measured <strong>by</strong> (a) zetasizer<br />

and (b) AFM. (From Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., J. Control Release, 90, 363, 2003.<br />

With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 449<br />

Transmittance %<br />

102<br />

100<br />

98<br />

96<br />

94<br />

92<br />

0<br />

6.0 6.5 7.0 7.5 8.0 1<br />

10<br />

100 1,000<br />

(a) pH<br />

(b)<br />

Particle size (nm)<br />

FIGURE 22.4 pH-dependent micelle destabilization. (a) The transmittance change of polyHis5K/PEG2K<br />

polymeric micelles (0.1 g/L) with pH. The polyHis5K/PEG2K polymeric micelles were prepared in<br />

NaOH–Na2B4O7 buffer solution (pH 8.0) and exposed at each pH for 24 h. (b) The change of particle size<br />

distribution in polyHis5K/PEG2K micelles exposed to each pH solution for 24 h; pH 8.0 (C), 7.4 (&), 7.2<br />

(:), 6.8 (;), 6.4 (%) and 6.0 (B). (From Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., J. Control Release,<br />

90, 363, 2003. With permission.)<br />

micellar structure of polyHis/PEG improved micelle stability at pH 7.4 and shifted destabilization<br />

to lower pHs. In particular, the mixed micelles composed of polyHis/PEG (75 wt%) and<br />

PLLA/PEG (25 wt%) showed a unimodal size distribution with an average size of 70-nm diameter<br />

at pH 9.0. Other mixed micelles prepared from different contents of PLLA/PEG ranged 70–100 nm<br />

in diameter with unimodal size distributions at pH 9.0. 41<br />

Figure 22.5 shows the effect of PLLA/PEG content on the CMC of the mixed micelles. When<br />

the mixed micelle was prepared at pH 9.0 and tested at pH 9.0, the CMC of the mixed micelles was<br />

very slightly influenced <strong>by</strong> PLLA/PEG content in a range of 0–25 wt%. However, above 25 wt%<br />

PLLA/PEG the CMC gradually increased, reaching to the higher CMC value of the PLLA/PEG<br />

micelle, although the CMC change is not statistically significant (Figure 22.5a). This indicates that<br />

a small amount of PLLA/PEG does not disturb the core stability.<br />

The presence of PLLA/PEG did significantly influence the pH-dependent CMC of the micelles.<br />

The CMC of polyHis/PEG micelle sharply increased from 2.7 to 140 mg/mL as pH decreased from<br />

8 to 5, with a major transition occurring between pH 7.0 and 6.8 (Figure 22.5b). However, the mixed<br />

micelle having 25 wt% PLLA/PEG showed a small increase in CMC, from 2.6 to 3.5 mg/mL, and<br />

the transition occurred in a pH range of 7.5–6.8 (Figure 22.5c).<br />

Figure 22.6 shows the pH-dependent stability of the mixed micelles monitored <strong>by</strong> relative light<br />

transmittance (T/T i%; the transmittance of the mixed micelle at each pH/transmittance at pH 9.0).<br />

The T/Ti% of polyHis/PEG micelle slightly increased with decreasing pH, especially below pH 7.6,<br />

which is attributed to micelle destabilization followed <strong>by</strong> dissociation. PLLA/PEG micelle, being<br />

pH insensitive, showed no change in turbidity and preserved its stability in the entire pH<br />

range tested.<br />

It is interesting to note that the T/Ti% of the mixed micelles decreased rather than increased with<br />

decreasing pH. This result seems to be associated with ionization of polyHis/PEG, accompanied <strong>by</strong><br />

separation and isolation of PLLA/PEG from the micelles, which in turn became segregated in<br />

water, and indirectly reflects the disintegration of the micelles <strong>by</strong> ionization of the imidazole<br />

groups along polyHis chains. The destabilizing pH was influenced <strong>by</strong> the amount of PLLA/PEG<br />

block copolymer in the mixed micelles. The addition of up to 10 wt% of PLLA/PEG to the micelle<br />

showed a slight shift in destabilizing pH. Although a 25 wt% addition considerably enhanced the<br />

micelle stability at pH 7.4 and destabilization only occurred below pH 7.0. At 40 wt% PLLA/PEG,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Intensity (%)<br />

40<br />

30<br />

20<br />

10


450<br />

CMC (μg/ml)<br />

4.0<br />

3.5<br />

3.0<br />

2.5<br />

2.0<br />

0<br />

20<br />

40 60 80 100<br />

(a) Adding content (wt%) of PLLA/PEG<br />

(b)<br />

T / Tinf %<br />

CMC (μg/mL)<br />

(c)<br />

110<br />

100<br />

90<br />

80<br />

3.8<br />

3.6<br />

3.4<br />

3.2<br />

3.0<br />

2.8<br />

2.6<br />

5.0<br />

5.5 6.0<br />

6.5 7.0 7.5 8.0<br />

5 6 7 8 9<br />

pH<br />

FIGURE 22.6 pH-dependent relative turbidity (T/Ti%) of the mixed micelles composed of polyHis/PEG and<br />

PLLA/PEG (PLLA/PEG content: 0 wt% (B); 5 wt% (C); 10 wt% (&); 25 wt% (:); 40 wt% (;); and<br />

100 wt% (,)). T i is the transmittance at 500 nm at pH 9.0 and T is the transmittance at a given pH after<br />

micelles were stabilized for 24 h. (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 91, 103, 2003.<br />

With permission.)<br />

CMC (μg/ml)<br />

pH<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Nanotechnology for Cancer Therapy<br />

0<br />

5.0 5.5 6.0 6.5 7.0 7.5 8.0<br />

FIGURE 22.5 The CMC change of the mixed micelles prepared with (a) different contents of PLLA/PEG in<br />

NaOH–Na2B4O7 pH 9.0 buffer, (b) pH-dependent CMC of polyHis/PEG micelle in two solutions (HCl or<br />

NaOH–Na 2B 4O 7 ionic strengthZ0.1), and (c) pH-dependent CMC of the mixed micelle composed of 25 wt%<br />

PLLA/PEG and 75 wt% polyHis/PEG in the same solutions (nZ3). (From Lee, E. S., Shin, H. J., Na, K., and<br />

Bae, Y. H., J. Control Release, 90, 363, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

pH


Polymeric Micelles Targeting Tumor pH 451<br />

the destabilization pH was shifted a bit further downward with less release of PLLA/PEG due to its<br />

high content.<br />

22.5 TUMOR pHe TARGETING<br />

The pH-dependent micelle property prompted us to test the micelles with pH-induced drug release.<br />

When doxorubicin (DOX) was incorporated into the mixed micelles (0, 10, 25, and 40 wt%<br />

PLLA/PEG) during dialysis, the micelle sizes ranged from 50 to 70 nm as determined <strong>by</strong><br />

dynamic light scattering. The drug loading efficiency was about 75–85% and the drug content in<br />

the micelles was 15–17 wt%. 41<br />

Figure 22.7a shows the cumulative DOX release from polyHis/PEG micelle. The DOX release<br />

pattern followed nearly first-order kinetics and reached a plateau in 24 h and 30 wt% of<br />

loaded DOX was released for 24 h at pH 8.0. Decreasing pH accelerated the release rate. The<br />

pH-dependent release patterns of the mixed micelles are presented in Figure 22.7b <strong>by</strong> plotting<br />

cumulative amount of DOX for 24 h vs. release medium pH. The mixed micelles containing<br />

25 wt% PLLA/PEG showed a favorable pH-dependency such that 32 wt% of DOX was released<br />

at pH 7.0, 70 wt% of DOX at pH 6.8, and 82 wt% at pH 5.0. On the other hand, the mixed micelles<br />

containing 40 wt% of PLLA/PEG suppressed the release of DOX to 35 wt% at pH 6.8, and 64 wt%<br />

at pH 6.6. This indicates that the content of PLLA/PEG controlled pH-dependent release from the<br />

mixed micelles <strong>by</strong> destabilization; the transition pH coincided with the results shown in<br />

Figure 22.6. The results support the idea that the mixed micelles truly recognize the minute<br />

difference in pH and discriminate the tumor pH <strong>by</strong> destabilization and release of micelle content.<br />

When the blank micelles were tested with human breast adenocarcinoma (MCF-7) cells, no<br />

noticeable cytotoxicity was observed up to 100 mg/mL of polymeric micelle for a 48-h culture,<br />

regardless of the culture medium pH. However, DOX-loaded micelles did present tumor cell<br />

toxicity in a pH-dependent manner.<br />

DOX-loaded polyHis/PEG micelles demonstrated a certain degree of cytotoxic effect at pH 7.4<br />

because of low micelle stability and DOX release. But the cell viability was much reduced at pH 6.8<br />

due to enhanced DOX release (Figure 22.8). The mixed micelles with 10 wt% PLLA/PEG<br />

(Figure 22.8b) caused low cytotoxicity at pH 7.4, which increases dramatically below pH 7.2.<br />

The above results are distinguished from the cytotoxicity of free DOX, which is almost independent<br />

of pH. The mixed micelles having 25 wt% PLLA/PEG (Figure 22.8c) appeared more sensitive to<br />

Cumulative DOX release (%)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Total DOX release (%) in 24 h<br />

0<br />

0 10 20 30 40 50<br />

20<br />

5.0 5.5 6.0 6.5 7.0 7.5 8.0<br />

(a) Time (h)<br />

(b)<br />

pH<br />

FIGURE 22.7 pH-dependent cumulative DOX release from the mixed micelles composed of polyHis/PEG<br />

and PLLA/PEG with varying PLLA/PEG content in the mixed micelles: (a) 0 wt%; pH 8.0 (C); 7.4 (&), 7.2<br />

(:), 7.0 (;), 6.8 (%), 6.2 (,), 5.0 (6). (b) 0 wt% (†), 10 wt% (&), 25 wt% (:) and 40 wt% (;) after 24 h.<br />

(From Lee, E. S, Na, K., and Bae, Y. H., J. Control Release, 91, 103, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

100<br />

80<br />

60<br />

40


452<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(a)<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(e)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1<br />

20<br />

1<br />

10 100 1,000 10,000<br />

DOX ng/mL<br />

(c) DOX ng/mL<br />

(d)<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

100<br />

80<br />

60<br />

40<br />

100<br />

80<br />

60<br />

40<br />

10 100 1,000 10,000<br />

20<br />

1 10 100<br />

DOX ng/mL<br />

1,000 10,000<br />

pH, differentiating especially between pH 7.0 and 6.8. The mixed micelles with 40 wt% PLLA/PEG<br />

micelles (Figure 22.8d) distinguished between pH 6.8 and 6.6 in a similar cytotoxicity study.<br />

Considering that a PLLA/PEG micelle (Figure 22.8e) has no pH-dependent cytotoxicity and<br />

neither do blank micelles, the pH-dependent cytotoxicity solely relies on the release of DOX at<br />

different pHs. The change in pH from 8.0 to 6.6 may influence cell surface charges, and cellular<br />

physiology and viability, noting that the low pH is a favorable environment for tumor cells but<br />

opposite for normal cells. The cell viability expressed in this study is relative to those at different<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(b)<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

100<br />

80<br />

60<br />

40<br />

20<br />

1 10 100 1,000 10,000<br />

100<br />

80<br />

60<br />

40<br />

Nanotechnology for Cancer Therapy<br />

DOX ng/mL<br />

20<br />

1 10 100 1,000 10,000<br />

DOX ng/mL<br />

FIGURE 22.8 The cytotoxicity of DOX-loaded mixed micelles with different PLLA/PEG contents (a) 0 wt%,<br />

(b) 10 wt%, (c) 25 wt%, (d) 40 wt%, and (e) 100 wt% after 48 h incubation at varying pH (pH 7.4 (C);<br />

7.2 (&); 7.0 (:); 6.8 (;); and 6.6 (%)). Free-DOX cytotoxicity is presented as a reference at pH 7.4 (B),<br />

7.2 (,), 7.0 (6), 6.8 (7), 6.6 (&). (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 91, 103, 2003.<br />

With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 453<br />

pHs in the absence of DOX, and the direct pH effect on the cell viability was not monitored with the<br />

MCF-7 cell line. No noticeable difference in cell viability <strong>by</strong> pH with free DOX was observed.<br />

In vitro results of the mixed micelles with 25 wt% PLLA/PEG (referred as PH-sensitive mixed<br />

micelle (PHSM) hereafter) showed the pH-dependent cytotoxicity of DOX against sensitive MCF-7<br />

cells, and were comparable with free DOX and a conventional pH-insensitive PLLA/PEG micelle<br />

(referred to as pH-insensitive micelle (PHIM). The free DOX showed the highest destruction, which<br />

is not significantly influenced <strong>by</strong> pH variation between 6.8 and 7.4. PHIM showed the least cytotoxicity<br />

at the two tested pH values. However, PHSM showed a distinct pH effect <strong>by</strong> switching the<br />

cytotoxicity from the level of PHIM at pH 7.0 (MCF-7 cell viability was 85% after the treatment of<br />

equivalent DOX 1 mg/mL) to a similar level of free DOX at pH 6.8 (MCF-7 cell viability was 36%<br />

after the treatment of equivalent DOX 1 mg/mL). This was attributed to pH-triggered release of DOX.<br />

Figure 22.9a shows the in vivo results of the anti-cancer activity of the tested DOX formulations<br />

i.v. injected on days 0 and 3. The DOX loaded PHSM (equivalent DOXZ10 mg/kg)<br />

exhibited significant inhibition (P!0.05 compared with free DOX or saline solution) of the<br />

growth of s.c. MCF-7 xenografts. The tumor volume of mice treated with the PHSM (P!0.05<br />

compared with free DOX) was approximately 4.5 and 3.6 times smaller than that of saline solution<br />

or free DOX treatment after 6 weeks, respectively. The triggered release of DOX <strong>by</strong> tumor pHe (!<br />

7.0), after accumulation of the micelle in the tumor sites via the enhanced permeation and retention<br />

(EPR) mechanism, 42,43 may present a more effective modality in tumor chemotherapy, providing<br />

higher local concentrations of the drug at tumor sites (targeted high-dose cancer therapy), while<br />

providing a minimal release of the drug from micelles while circulating in the blood (pH 7.4). When<br />

compared with PHIM, the volume of tumors treated with the PHSM was approximately 3.0 times<br />

smaller (particle sizeZ50–70 nm, data now shown) than those whose properties are not influenced<br />

<strong>by</strong> pH. In addition, the destabilization of PHSM (converting to unimers) may help extravasation of<br />

additional micelles from the blood compartment and their penetration into tumors <strong>by</strong> reducing the<br />

physical barrier effect, which can be generated <strong>by</strong> stable particles, such as conventional liposomes<br />

and polymeric micelles. It was reported that drug carriers, such as PEG modified liposomes, are in<br />

an insoluble form and only reside in the vicinity of the leaky blood vessels after extravasation,<br />

rather than migrating into the deeper sites of tumors. This “road block effect” could obstruct the<br />

subsequent accumulation of the carriers at tumor sites. 44 It is interesting to note that only PHSM<br />

administration showed decreased tumor size for the initial first week. No obvious changes in mice<br />

Tumor volume (mm 3 )<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

14<br />

0 5 10 15 20 25 30 35 40 45 0 10 20 30 40<br />

(a) Time (day)<br />

(b)<br />

Time (day)<br />

FIGURE 22.9 Tumor growth inhibition of s.c. human breast MCF-7 carcinoma xenografts in BALB/c nude<br />

mice. Mice were injected i.v. with 10 mg/kg DOX equivalent dose: PHSM (C), PHIM (&), free DOX (:),<br />

and saline solution (;). Two i.v. injections on day 0 and 3 were administered. (a) Tumor volume change and<br />

(b) body weight change (nZ5). Values are the meansGthe standard deviations (SD). (From Lee, E. S., Na, K.,<br />

and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Body weight (g)<br />

26<br />

24<br />

22<br />

20<br />

18<br />

16


454<br />

body weight in experimental groups, except for the free DOX group, were observed. Free DOX<br />

caused more weight-loss after i.v. administration (Figure 22.9b) than micelle formulations.<br />

“Stealth” particles decorated with PEG are known to be invisible to macropharges and have<br />

prolonged half-lives in the blood compartment. This property endows the particles a higher probability<br />

to extravasate in pathological sites, like tumors or inflamed regions where a leaky vascular<br />

structure persists. 45 Recent investigations have reported that the size cut-off of tumor vasculature for<br />

the EPR effect grew to 200–700 nm. 46 Nevertheless, the nanoparticle penetration to tumor vasculature<br />

is heterogeneous in a solid tumor, and the vascular structure or cut off size differs from tumor<br />

to tumor and is influenced <strong>by</strong> tumor location. 46 A more recent study demonstrated that the optimal<br />

size of stealth liposomes was determined to be around 100 nm in an animal model. 47 This means that<br />

the size of micelle or liposome affects the in vivo bio-distribution, post injection. Smaller size causes<br />

more accumulation in the liver, while larger ones cause more accumulation in the spleen. 48<br />

We hypothesized that the average particle size of 50–70 nm of PHSM may meet the optimal<br />

size for extravasation or long-circulation suggested <strong>by</strong> several research groups. 42,44,47,48 A significant<br />

portion of DOX carried <strong>by</strong> the PHSM, however, was accumulated in the liver and spleen. In<br />

Figure 22.10a, DOX levels in the blood after i.v. administration of PHSM to the animals bearing s.c.<br />

MCF-7 carcinoma xenografts were 27% at 4 h, 8% at 12 h, 4% at 48 h in the blood and 18% at 4 h,<br />

15% at 12 h, 3% at 48 h in the tumors and 14% at 4 h, 7% at 12 h, 1.5% at 48 h in the liver and 2% at<br />

4 h, 1% at 12 h, barely detectable levels at 48 h in the spleen. The mononuclear phagocytes system<br />

(MPS), especially in the liver and spleen, 42,44,46,49 seems to be responsible for the accumulation in<br />

these organs even with small size of the particles.<br />

As shown in Figure 22.8b, because of very short half-life (t1/2, within few minutes) of free DOX<br />

in the blood, 50 DOX levels in the blood at 12–48 h were not detectable. DOX levels in the tumors<br />

were 1.5% at 4 h, 1% at 12 h, and barely detectable levels were found at 48 h. In contrast, high<br />

concentrations of DOX were found in the heart, liver and considerable DOX levels were shown in<br />

Total DOX (%) in orans<br />

50<br />

40<br />

30<br />

20<br />

10<br />

(a) 0<br />

40<br />

30<br />

20<br />

10<br />

(b) 0<br />

tumor blood liver heart spleen lung kidney<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 22.10 Total DOX (%) levels (total DOX remaining in organs to DOX amount first injected) in blood<br />

and tissues (tumor, liver, heart, spleen, lung, kidney) after i.v. administration (10 mg/kg DOX equivalent dose)<br />

of (a) PHSM micelles, and (b) free DOX. The mice (s.c. human breast MCF-7 carcinoma xenografts) were<br />

sacrificed at 4 (&), 12 (&), and 48 h (&) after i.v. administration. Values are the meansGthe standard<br />

deviations (nZ4, SD). (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 455<br />

the lung, spleen and kidney. These results are in striking contrast to DOX biodistribution when<br />

carried <strong>by</strong> PHSM. Especially, PHSM is responsible for the enhanced tumor accumulation. 51 The<br />

combined mechanisms of pH e-triggered release and EPR effect can be beneficial to treat solid<br />

tumors <strong>by</strong> minimizing the road block effect of the particles after extravasation, increasing local<br />

concentration for diffusion to the tumors and more active internalization. This, however, requires<br />

further investigation for verification.<br />

22.6 EARLY ENDOSOMAL pH TARGETING FOR REVERSAL OF MDR<br />

PolyHis has been known for endosomolytic activity and recently low MW polyHis-conjugated<br />

poly(L-lysine) has demonstrated improved gene transfection. 52 However, the difficulty in controlling<br />

its molecular weight and limited solubility of high MW polymer in organic solvents has restricted the<br />

fabrication and applications of polyHis-based drug carriers. The mixed micelles, who’s surface was<br />

decorated with folate, consisting of polyHis/PEG-folate (Figure 22.11a) and PLLA/PEG-folate<br />

(Figure 22.11b), showed a similar release pattern of the micelles without folate (PHSM); cumulative<br />

DOX release for 24 h of 30–35 wt% in the pH range of 7.0–7.4, which is slightly more than 20–<br />

30 wt% initial burst release from conventional pH-insensitive micelles, 53 but about 70 wt% at pH<br />

6.8 during the same time period. During this process, it was observed that the drug or polymer was<br />

distributed in the cytosol and the nucleus in a sensitive breast cancer cell line (MCF-7). 41<br />

From these results, we hypothesized that the folate receptor-mediated internalization route and<br />

the micelle destabilization reverse the drug resistance of solid tumors. For proof of concept, a MCF-<br />

7/DOX R cell was selected after being stepwise exposed to 0.001–10 mg/mL of free DOX as<br />

reported. 54–58<br />

Figure 22.12 shows the decreased cytotoxicity of free DOX against MCF-7/DOX R cells, exhibiting<br />

a typical drug resistance curve. The cytotoxicity of DOX against sensitive MCF-7 cells<br />

slightly decreased <strong>by</strong> lowering pH from 8.0 to 6.8 because of an increased degree of ionization<br />

of DOX, causing less diffusivity through the plasma membrane. 59 For the resistant cells, the pH<br />

effect was not noticed even at a high DOX concentration range. The IC50 value of free DOX<br />

increased from 470 ng/mL for wild MCF-7 cells to 25,800 ng/mL for MCF-7/DOX R cells at pH 7.4.<br />

Figure 22.13 visualizes different distributions of free DOX in MCF-7 (Figure 22.13a) and<br />

MCF-7/DOX R (Figure 22.13b) cells. In the MCF-7/DOX R cells, there was a limited DOX uptake<br />

in the cytosolic compartment, while the sensitive MCF-7 cells accumulated DOX in the cytosolic<br />

compartment and nucleus. Figure 22.14 shows the cytotoxic effects of free DOX or DOX-loaded<br />

(a)<br />

(b) H2N HN<br />

H 2 N<br />

HN<br />

O<br />

N<br />

O<br />

N<br />

N<br />

N<br />

N N<br />

O<br />

N<br />

H N<br />

H<br />

COOH<br />

O<br />

O COOH<br />

N<br />

H N<br />

H<br />

O<br />

O<br />

PEG C<br />

H<br />

O<br />

N H C C<br />

CH 2<br />

N N H<br />

FIGURE 22.11 The chemical structures of (a) polyHis/PEG-folate and (b) PLLA/PEG-folate.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

H<br />

N<br />

N<br />

H<br />

O<br />

C<br />

PEG<br />

O<br />

O<br />

H CH3 N CH<br />

n<br />

CH 3<br />

CH 3<br />

O<br />

n<br />

H


456<br />

Cell Viability (%) <strong>by</strong> MTT assay<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1 10 100 1000 10000 100000<br />

DOX ng/mL<br />

FIGURE 22.12 Characterization of micelles on MCF-7/DOX R cells: The cytotoxicity of free DOX against<br />

MCF-7 (close circle) and MCF-7/DOX R (open circle) cells at pH 8.0 (C), pH 7.4 (&) and pH 6.8 (:)(nZ7).<br />

(From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

FIGURE 22.13 Confocal microscopy studies on (a) DOX-sensitive MCF-7 and (b) DOX-resistant MCF-7<br />

(MCF-7/DOX R ) cells treated with free DOX at pH 6.8.<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(a)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0 0 10 100 1000 10000<br />

DOX ng/mL<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(b)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Nanotechnology for Cancer Therapy<br />

0<br />

1 10 100 1000 10000<br />

DOX ng/mL<br />

FIGURE 22.14 The cytotoxicity of DOX-loaded PHSM/f (C), polyHis/PEG-folate micelles (&),<br />

PLLA/PEG-folate micelles (:), free DOX (;) against MCF-7/DOX R cells at (a) pH 7.4 and (b) pH 6.8<br />

after 48 h incubation (nZ7). (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 457<br />

micelles against MCF-7/DOX R cells at pH 7.4 and 6.8. The viability of MCF-7/DOX R cells was<br />

not dramatically decreased as expected at even high concentrations of free DOX due to the<br />

influence of Pgp function. DOX-loaded PLLA/PEG-folate micelles (pH-insentive micelles as a<br />

control) showed limited cytotoxicity against MCF-7/DOX R cells. DOX loaded polyHis/PEGfolate<br />

micelles demonstrated a similar degree of cytotoxicity against MDR cells as that of free<br />

DOX against the sensitive cells. 41 This effect is most likely due to enhanced DOX release and the<br />

endosomal escaping activity of polyHis/PEG-folate unimer released after internalization of the<br />

micelles and subsequent destabilization. This can be regarded as cytosolic delivery of DOX,<br />

disturbing drug sequestration mechanism of MDR. 41 However, it is shown that the viability of<br />

cells treated <strong>by</strong> PHSM with folate (PHSM/f) was about 10% at 10 mg/mL (DOX concentration),<br />

whereas for polyHis/PEG-folate micelles (without PLLA/PEG-folate) the value is about 40% at<br />

similar DOX concentration (Figure 22.14a). This enhanced effect suggests that there could be an<br />

additional mechanism of PHSM/f beside the functions of polyHis. Hoffman’s group reported that<br />

the membrane disruption mechanisms of poly(2-ethylacrylic acid) (PEAA) is presumably related<br />

to their increase in hydrophobicity at low pH values. 60 Similar to that, PLLA/PEG-folate released<br />

from destabilized PHSM/f could interact with the endosomal membrane due to the exposed<br />

hydrophobic PLLA portion, disturbing membrane stability. There is no significant difference<br />

between two pHs tested (pH 7.4 (Figure 22.14a) and 6.8 (Figure 22.14b)). Figure 22.14b suggests<br />

the feasibility of MDR reversal <strong>by</strong> PHSM/f at acidic tumor pH e. 61<br />

In the case of sensitive MCF-7 cells, PHSM/f showed significant cytotoxicity against cells.<br />

Figure 22.15 demonstrates this effect. The DOX-loaded micelles with folate were internalized via<br />

folate-receptor mediated endocytosis. The PHSM/f micelles induced greatly enhanced cytotoxicity<br />

(cell viability 16% at DOX 10 mg/mL at pH 8.0 (this pH was employed to ensure the micelle<br />

stability before internalization)) resulting from increased micelle concentration inside the cell due<br />

to folate receptor-mediated endocytosis and fusogenic effect of polyHis, while the PLLA/PEGfolate<br />

showed a slight increase in cytotoxicity. These observations clarify the role of polyHis in<br />

anti-cancer activity as two folds: (1) destabilization for drug release and (2) fusogenicity.<br />

Figure 22.15b indicates that, once internalization <strong>by</strong> folate receptors grew predominant, the pH<br />

effect on drug release became insignificant. This observation can be explained <strong>by</strong> the facts that (1)<br />

the mixed micelles were internalized into tumor cells via folate-receptor mediated endocytosis and<br />

underwent destabilization in a slightly acidic early endosomal compartment <strong>by</strong> protonation of an<br />

imidazole group and (2) DOX released from the mixed micelles under the influence of the extracellular<br />

pH, regardless of folate-receptor mediated endocytosis, has effective an drug portion for<br />

cell killing. Figure 22.15c shows the cell killing rates of the mixed micelles and free DOX. The free<br />

DOX diffuses only passively to cells. Unlike free DOX, PHSM/f facilitated faster cell killing in<br />

0–4 h of incubation. This observation is consistent with the high number of folate-receptors on<br />

tumor cells. In addition, there is little difference in the cell killing rate between pH 6.8 and 7.4 (data<br />

not shown), thus explaining quick cytotoxic action of DOX-loaded mixed micelles against MCF-7<br />

cells due to the active internalization. 58,59<br />

For further evidence of the functioning of PHSM/f, a confocal-microscopic study was<br />

conducted to reveal the distribution of DOX in the cells. Figure 22.16a shows the localized<br />

DOX in MCF-7/DOX R cells treated with DOX-loaded PLLA/PEG-folate micelles. The PLLA/<br />

PEG-folate micelles were entrapped in endosome or multivesicular bodies resulting from folate<br />

receptor-mediated endocytosis. This could be responsible for the discrete DOX-intensity and biased<br />

DOX-intensity distribution. The entrapment seems to be a consequence of the absence of fusogenic<br />

activity. Unlike PLLA/PEG-folate micelles, PHSM/f incubation resulted in broad distribution of<br />

DOX in cytosol as well as nuclear compartments (Figure 22.16b).<br />

To <strong>by</strong>pass one of the major MDR mechanisms, Pgp pumping, we employed folate receptormediated<br />

endocytosis, because MCF-7 cells over-express folate receptors. 62 Along with this entry<br />

mechanism, the further enhanced drug release at endosomal pH 36,41 and the disruption of the<br />

endosomal membrane 41,51,52,63 are expected to kill MDR cells.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


458<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

1 10 100 1000 10000<br />

(a) DOX ng/mL<br />

(b)<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

(c)<br />

110<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

0<br />

1 10<br />

0 5 10 15 20 25<br />

Incubation time (h)<br />

FIGURE 22.16 Confocal microscopy studies on MCF-7/DOX R cells treated with (a) DOX-loaded<br />

PLLA/PEG-folate micelles and (b) DOX-loaded PHSM/f micelles at pH 6.8. The cells were incubated with<br />

the carriers for 30 min. (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

Cell viability (%) <strong>by</strong> MTT assay<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Nanotechnology for Cancer Therapy<br />

100<br />

DOX ng/mL<br />

1000 10000<br />

FIGURE 22.15 The cytotoxicity of (a) free DOX (C), polyHis/PEG-folate micelle (&), PLLA/PEG-folate<br />

micelle (:) and PHSM/f (;) at pH 8.0 and of (b) PHSM/f at pH 7.4 (C) and pH 6.8 (&) against sensitive<br />

MCF-7 cells after 48 h incubation. (c) The cytotoxicity against MCF-7 cells treated with PHSM/f (C) and free<br />

DOX (&) at pH 6.8. The DOX content in the micelles was adjusted to be equivalent to free DOX concentration<br />

(500 ng/mL) and MCF-7 cells were treated for a given incubation time. (From Lee, E. S., Na, K., and Bae, Y.<br />

H., J. Control Release, 91, 103, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 459<br />

In the in vitro studies with MCF-7/DOX R cells, 51 the PHSM/f showed a significantly enhanced<br />

efficacy in tumor cell cytotoxicity, most likely due to enhanced internalization (folate receptormediated<br />

endocytosis), and subsequent pH-induced micelle destabilization and endosomal escape.<br />

However, the PHSM or PHIM with folate (PHIM/f) showed low efficacy in killing MCF-7/DOX R<br />

cells, due to lack of entry mechanism and/or endosomal escape.<br />

The MCF-7/DOX R xenografts in nude mice were used for the investigation of in vivo efficacy.<br />

The tumor bearing animals were treated <strong>by</strong> multiple i.v. injections on days 0, 3, and 6<br />

(Figure 22.17a). Neither complete tumor regression, nor toxicity-induced death was observed in<br />

all experimental groups, though, the tumor growth in mice treated <strong>by</strong> PHSM/f was inhibited and the<br />

size was reduced for two weeks after the third i.v. injection. The tumor volume in mice treated <strong>by</strong><br />

PHSM/f (P!0.05 compared with free DOX) was approximately 2.7-times smaller than those<br />

treated with free DOX or PHIM, and approximately 1.9-times smaller than those treated with<br />

PHSM after six weeks. These results indicate that the folate-mediated endocytosis pathway of<br />

PHSM/f enhances the regression of tumors and is more efficient for MDR tumor chemotherapy<br />

than PHSM. This being said, the tumor regression efficacy in PHIM/f was not significant because<br />

some folate conjugates may have recycled back to the surface 64 before drug unloading. In addition,<br />

slow or limited drug release from PHIM maybe associated with drug sequestration in acidic<br />

organelles in MDR cells. 31,33 The active internalization, enhanced drug release rate, and endosomal<br />

escaping activity justify the efficacy of PHSM/f. Free DOX administration exhibited more weightloss<br />

in mice than the groups treated <strong>by</strong> any micelles (Figure 22.17b), signifying lower toxicity of<br />

DOX-loaded micelles.<br />

As shown in Figure 22.18, s.c. MCF-7/DOX R carcinoma xenografts showed different DOX<br />

levels in tumor as compared to s.c. MCF-7 carcinoma xenografts, particularly for PHSM. DOX<br />

levels in the tumors of s.c. MCF-7/DOX R carcinoma xenografts in which the DOX levels were 21%<br />

at 4 h, 12% at 12 h, and 5% at 48 h for PHSM/f; 7% at 4 h, 5% at 12 h, and 2% at 48 h for PHSM;<br />

and 1% at 4 h and barely detectable levels at 12 and 48 h for free DOX. This suggests that the role of<br />

the Pgp pump in s.c. MCF-7/DOX R carcinoma xenografts is significant in determining DOX levels<br />

in the resistant tumors. In the case of DOX delivered <strong>by</strong> PHSM into the resistant tumor, the released<br />

DOX cannot diffuse in the cells but is instead being washed away from the site. On the other hand,<br />

DOX carried <strong>by</strong> PHSM/f shows similar accumulation levels in the resistant tumor to that of<br />

sensitive tumor, resulting from active internalization <strong>by</strong> folate receptor-mediated endocytosis.<br />

Tumor volume (mm 3 )<br />

1,000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 5 10 15 20 25 30 35 40 45<br />

14<br />

0 10 20 30 40<br />

(a) Time (day)<br />

(b)<br />

Time (day)<br />

FIGURE 22.17 Tumor growth inhibition of s.c. human breast MCF-7/DOX R carcinoma xenografts in<br />

BALB/c nude mice. Mice were injected i.v. with 10 mg/kg DOX equivalent dose: PHSM/f (C), PHSM<br />

(&), PHIM/f (:), PHIM (;), and free DOX (%). Three i.v. injections on day 0, 3, and 6 were made. (a)<br />

Tumor volume change and (b) body-weight change (nZ5). Values are the meansGthe standard deviations<br />

(SD). (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103, 405, 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Body weight (g)<br />

26<br />

24<br />

22<br />

20<br />

18<br />

16


460<br />

Total DOX (%) in organs<br />

This, combined with the endosomal escaping capacity of PHSM/f, resulted in a remarkable<br />

regression effect for resistant tumors, as shown in Figure 22.18.<br />

22.7 CONCLUSIONS<br />

40<br />

30<br />

20<br />

10<br />

(a) 0<br />

40<br />

30<br />

20<br />

10<br />

(b) 0<br />

40<br />

(c)<br />

30<br />

20<br />

10<br />

0<br />

Tumor blood liver heart spleen lung kidney<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 22.18 Total DOX (%) levels (total DOX remaining in organs to DOX content first injected) in blood<br />

and tissues (tumor, liver, heart, spleen, lung, and kidney) after i.v. administration (10 mg/kg DOX equivalent<br />

dose) of (a) PHSM/f micelles, (b) PHSM micelles, and (c) free DOX. Mice (s.c. human breast MCF-7/DOX R<br />

carcinoma xenografts) were sacrificed at 4 (&), 12 (&), and 48 h (&) after i.v. administration. Values are the<br />

meansGthe standard deviations (nZ4, SD). (From Lee, E. S., Na, K., and Bae, Y. H., J. Control Release, 103,<br />

405, 2005.)<br />

Although important findings in scientific research and technological advances, such as long-circulating<br />

carries, EPR effect, tumor specific receptor-mediated targeting, and MDR modulator and<br />

reverters, have been achieved in the last few decades for effective solid tumor targeting, current<br />

chemotherapy is still facing major challenges to improving drug accumulation in tumor sites and<br />

reversing intrinsic or acquired drug-resistance of tumors.<br />

An accelerated release rate of doxorubicin at tumor extracellular pH was provided <strong>by</strong> pH-sensitive<br />

polyHis/PEG micelles, but the micelle was not fully stable at pH 7.4. When a pH-insensitive<br />

block copolymer, PLLA–PEG, was blended with polyHis-PEG to form a mixed micelle, the<br />

stability of the micelles at pH 7.4 was significantly improved and the micelles’ destabilization<br />

pH shifted from 7.2 to 6.6 with an increase in the amount of PLLA–PEG in the mixed micelle.<br />

The micelles were further modified with folate. These pH-sensitive micelles proved to be much<br />

more effective in regressing wild-type MCF-7 cells in vitro and in vivo studies with less weight-loss<br />

in animals, when compared with the pH-insensitive PLLA–PEG micelle. The folate effect was not<br />

apparent with the sensitive cells in the in vitro results. It seems that doxorubicin released either<br />

extracellularly or intracellularly is effective for cell-kill due to its high diffusivity through the<br />

plasma membrane.<br />

The pH-sensitive micelle-with-folate showed a dramatic effect when tested with drug-resistant<br />

tumors in vitro and in vivo, demonstrating a similar cell-kill rate as that of sensitive cells.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 461<br />

The underlying mechanisms involve (1) fast internalization of the micelles to <strong>by</strong>pass the Pgp, (2)<br />

fast release rate of doxorubicin at early endosomes, and (3) quick disruption of the endosomal<br />

membrane <strong>by</strong> virtue of polyHis protonation and possible membrane interactions of freed polyHis-<br />

PEG. These mechanisms avoid the major MDR mechanisms of the Pgp pump and drug sequestration<br />

in acidic organelles and endosomal recycling, which resulted in uniform intracellular<br />

distributions of doxorubicin and polyHis-PEG, including the nucleus, In a resistant tumor xenograft<br />

model in mice, the tumor was regressed <strong>by</strong> iv administration of the folate-decorated pH-sensitive<br />

micelles with less weight-loss of the treated animal than control animals treated <strong>by</strong> pH-insensitive<br />

micelles and free doxorubicin.<br />

REFERENCES<br />

1. Levasseur, L. M., Slocum, H. K., Rustum, Y. M., and Greco, W. R., Modeling of the time-dependency<br />

of in vitro drug cytotoxicity and resistance, Cancer Res., 58, 5749, 1998.<br />

2. Gardner, S. N., A mechanistic, predictive model of dose–response curves for cell cycle phase-specific<br />

and -nonspecific drugs, Cancer Res., 60, 1417, 2000.<br />

3. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Long-circulating and target-specific nanoparticles:<br />

Theory and practice, Pharmacol. Rev., 53, 283, 2001.<br />

4. Yokoyama, M., Novel passive targetable drug delivery with polymeric micelles, In Biorelated Polymers<br />

and Gels, Okano, T., Ed., Academic Press, San Diego, pp. 193–229, 1998.<br />

5. Drummond, D. C., Meyer, O., Hong, K., Kirpotin, D. B., and Papahadjppoulos, D., Optimizing<br />

liposomes for delivery of chemotherapeutic agents to solid tumors, Pharmacol. Rev., 51, 691, 1999.<br />

6. Kong, G., Braun, R. D., and Dewhirst, M. W., Hyperthermia enables tumor-specific nanoparticle<br />

delivery: Effect of particle size, Cancer Res., 60, 4440, 2000.<br />

7. Munshi, N., Rapoport, N., and Pitt, W. G., Ultrasonic activated drug delivery from Pluronic P-105<br />

micelles, Cancer Lett., 118, 13, 1997.<br />

8. Hobbs, S. K., Monsky, W. L., Yuan, F., Roberts, W. G., Griffith, L., Torchilin, V. P., and Jain, R. K.,<br />

Regulation of transport pathways in tumor vessels: Role of tumor type and microenvironment, Proc.<br />

Natl Acad. Sci. USA, 95, 4607, 1998.<br />

9. Tannock, I. F. and Rotin, D., Acid pH in tumors and its potential for therapeutic exploitation, Cancer<br />

Res., 49, 4373, 1989.<br />

10. Stubbs, M., McSheehy, R. M. J., Griffiths, J. R., and Bashford, L., Causes and consequences of tumour<br />

acidity and implications for treatment, Opinion, 6, 15, 2000.<br />

11. Engin, K., Leeper, D. B., Cater, J. R., Thistlethwaite, A. J., Tupchong, L., and McFarlane, J. D.,<br />

Extracellular pH distribution in human tumors, Int. J. Hyperthermia, 11, 211, 1995.<br />

12. van Sluis, R., Bhujwalla, Z. M., Raghunand, N., Ballesteros, P., Alvarez, J., Cerdán, S., Galons, J. P.,<br />

and Gillies, R. J., In vivo imaging of extracellular pH using 1 H MRSI, Magn. Reson. Med., 41, 743,<br />

1999.<br />

13. Ojugo, A. S. E., McSheehy, P. M. J., Mcintyre, D. J. O., McCoy, C., Stubbs, M., Leach, M. O., Judson,<br />

I. R., and Griffiths, J. R., Measurement of the extraceullar pH of solid tumours in mice <strong>by</strong> magnetic<br />

resonace spectroscopy: A comparison of exogenous 19 F and 31 P probes, NMR Biomed., 12, 495, 1999.<br />

14. Leeper, D. B., Engin, K., Thistlethwaite, A. J., Hitchon, H. D., Dover, J. D., Li, D. J., and Tupchong,<br />

L., Humon tumor extracellular pH as a function of blood glucose concentration, Int. J. Radiat. Oncol.<br />

Biol. Phys., 28, 935, 1994.<br />

15. Yamagata, M., Hasuda, K., Stamato, T., and Tannock, I. F., The contribution of lactic acid to<br />

acidification of tumours: Studies of variant cells lacking lactate dehydrogenase, Br. J. Cancer, 77,<br />

1726, 1998.<br />

16. Gerasimov, O. V., Boomer, J. A., Qualls, M. M., and Thompson, D. H., Cytosolic drug delivery using<br />

pH- and light-sensitive liposomes, Adv. Drug Deliv. Rev., 38, 317, 1999.<br />

17. Drummond, D. C., Zignani, M., and Leroux, J. C., Current status of pH-sensitive liposomes in drug<br />

delivery, Prog. Lipid Res., 39, 409, 2000.<br />

18. Gottesman, M. M., Pastan, I., and Ambudkar, S. V., P-glycoprotein and multi drug resistance, Curr.<br />

Opin. Genet. Dev., 6, 610, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


462<br />

Nanotechnology for Cancer Therapy<br />

19. Narasaki, F., Matsuo, I., Ikuno, N., Fukuda, M., Soda, H., and Oka, M., Multidrug resistance-associated<br />

protein (MRP) gene expression in human lung cancer, Anticancer Res., 16, 2079, 1996.<br />

20. Zaman, G. J., Lankelma, J., van Tellingen, Q., Beijnen, J., Dekker, H., Paulusma, C., Oude Elferink,<br />

R. P., Baas, F., and Borst, P., Role of gluthathione in the export of compounds from cells <strong>by</strong> the<br />

multidrug resistance-associated protein, Proc. Natl Acad. Sci. USA, 92, 7690, 1995.<br />

21. De Jong, S., Zijlstra, J. G., de-Vries, E. G., and Mulder, N. H., Reduced DNA topoisomerase II activity<br />

and drug-induced DNA cleavage activity in an adriamycin-resistant human small cell lung carcinoma<br />

cell line, Cancer Res., 50, 304, 1990.<br />

22. Sugawara, I., Akiyama, S., Scheper, R. J., and Itoyama, S., Lung resistance protein (LRP) expression<br />

in human normal tissues in comparison with that of MDR1 and MRP, Cancer Lett., 112, 23, 1997.<br />

23. Pohl, G., Filipits, M., Suchomel, R. W., Stranzl, T., Depisch, D., and Pirker, R., Expression of the lung<br />

resistance protein (LRP) in primary breast cancer, Anticancer Res., 19, 5051, 1999.<br />

24. William, S. D. and Scheper, R. J., Lung resistance-related protein: Determining its role in multidrug<br />

resistance, J. Natl Cancer Inst., 91, 1604, 1999.<br />

25. Gonçlaves, A., Braguer, D., Kamath, K., Martello, L., Briand, C., Horwitz, S., Wilson, L., and Jordan,<br />

M. A., Resistance to taxol in lung cancer cells associated with increased microtubule dynamics, Proc.<br />

Natl. Acad. Sci. USA, 98, 11737, 2001.<br />

26. Rogan, A. M., Hamilton, T. C., Young, R. C., Klecker, R. W., and Ozols, R. F., Reversal of adriamycin<br />

resistance <strong>by</strong> verapamil in human ovarian cancer, Science, 224, 994, 1984.<br />

27. Germann, U. A., Shlyakhter, D., Mason, V. S., Zelle, R. E., Duffy, J. P., Gallulo, U., Armistead, D. M.,<br />

Saunders, J. O., Boger, J., and Harding, M. W., Cellular and biochemical characterization of VX-710<br />

as a chemosensitizer: Reversal of P-glycoprotein-mediated multidrug resistance in vitro, Anticancer<br />

Drugs, 8, 125, 1997.<br />

28. Dale, I. L., Tuffley, W., Callaghan, R., Holmes, J. A., Martin, K., Luscombe, M., Mistry, P., Ryder,<br />

H., Stewart, A. J., Charlton, P., Twentyman, P. R., and Bevan, P., Reversal of P-glycoproteinmediated<br />

multidrug resistance <strong>by</strong> XR9051, a novel diketopiperazine derivative, Br. J. Cancer, 78,<br />

885, 1998.<br />

29. Fisher, G. A., Lum, B. L., Hausdorff, J., and Sikic, B.I, Pharmacological considerations in the<br />

modulation of multidrug resistance, Eur. J. Cancer, 32A, 1082, 1996.<br />

30. Mistry, P., Stewart, A. J., Dangerfield, W., Okiji, S., Liddle, C., Bootle, D., Plumb, J. A., Templetion,<br />

D., and Charlton, P., In vitro and in vivo reversal of P-glycoprotein-mediated multidrug resistance <strong>by</strong><br />

a noel potent modulator, XR9576, Cancer Res., 61, 749, 2001.<br />

31. Simon, S. M., Role of organelle pH in tumor cell biology and drug resistance, Drug Discov. Today,4,<br />

32, 1999.<br />

32. Busa, W. B. and Nuccitelli, R., Metabolic regulatoin via intracellular pH, Am. J. Physiol., 246, R409,<br />

1984.<br />

33. Simon, S., Roy, D., and Schindler, M., Intracellular pH and the control of multidrug resistance, Proc.<br />

Natl Acad. Sci. USA, 91, 1128, 1994.<br />

34. Belhoussine, R., Morjani, H., Millot, J. M., Sharonov, S., and Manfait, M., Confocal scanning<br />

microspectrofluorometry reveals specific anthracycline accumulation in cytoplasmic organelles of<br />

multidrug-resistant cancer cells, J. Histochem. Cytochem., 46, 1369, 1998.<br />

35. Tang, F., Horie, K., and Borchardt, R. T., Are MDCK cells transfected with the human MDR1 gene a<br />

good model of the human intestinal mucosa?, Pharm. Res., 19, 765, 2002.<br />

36. Lee, E. S., Shin, H. J., Na, K., and Bae, Y. H., Poly(L-histidine)-PEG block copolymer micelles and<br />

pH-induced destabilization, J. Control Release, 90, 363, 2003.<br />

37. Patchornik, A., Berger, A., and Katchalski, E., Poly-L-histidine, J. Am. Chem. Soc., 79, 5227, 1957.<br />

38. Urry, D. W., Peng, S., Gowda, D. C., Parker, T. M., and Harris, R. D., Comparison of electrostaticand<br />

hydrophobic-induced pKa shifts in polypentapeptides. The lysine residue, Chem. Phys. Lett., 225,<br />

97, 1994.<br />

39. Marques, C. M., Bunchy micelles, Langmuir, 13, 1430, 1997.<br />

40. La, S. B., Okano, T., and Kataoka, K., Preparation and characterization of the micelle-forming polymeric<br />

drug indomethacin-incorporated poly(ethylene oxide)-poly(beta-benyzl-L-aspartate) block<br />

copolymer micelle, J. Pharm. Sci., 85, 85, 1996.<br />

41. Lee, E. S., Na, K., and Bae, Y. H., Polymeric micelle for tumor pH and folate mediated targeting,<br />

J. Control Release, 91, 103, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Polymeric Micelles Targeting Tumor pH 463<br />

42. Ning, Z., Da, D., Rudoll, T. L., Needham, D., Whorton, A. R., and Dewhirst, M. W., Increased<br />

microvascular permeability contributes to preferential accumulation of stealth liposomes in tumor<br />

tissue, Cancer Res., 53, 3764, 1993.<br />

43. Maeda, H., Wu, J., Sawa, T., Matsumura, Y., and Hori, K., Tumor vascular permeability and the EPR<br />

effect in macromolecular therapeutics: A review, J. Control Release, 65, 271, 2000.<br />

44. Yuan, F., Leunig, M., Huang, S. K., Berk, D. A., Papahadjopoulos, D., and Jain, R. K., Microvascular<br />

permeability and interstitial penetration of sterically stabilized (stealth) liposomes in a human tumor<br />

xenograft, Cancer Res., 54, 4564, 1994.<br />

45. Oku, N., Namba, Y., and Okada, Y., Tumor accumulation of novel RES-avoiding liposomes, Biochim.<br />

Biophys. Acta., 1126, 225, 1992.<br />

46. Hobbs, S. K., Monsky, W. L., Yuan, F., Roberts, W. G., Griffith, L., Torchilin, V. P., and Jain, R. K.,<br />

Regulation of transport pathways in tumor vessels: Role of tumor type and microenvironment, Proc.<br />

Natl Acad. Sci. USA, 95, 4607, 1998.<br />

47. Charrois, G. J. R. and Allen, T. M., Rate of biodistribution of STEALTH w liposomes to tumor and<br />

skin: Influence diameter and implications for toxicity and therapeutic activity, Biochim. Biophys.<br />

Acta., 1609, 102, 2003.<br />

48. Takeuchi, H., Kojima, H., Toyoda, T., Yamamoto, H., Hinox, T., and Kawashima, Y., Prolonged<br />

circulation time of doxorubicin-loaded liposomes coated with a modified polyvinyl alcohol after<br />

intravenous injection in rats, Eur. J. Pharm. Biopharm., 48, 123, 1999.<br />

49. Goren, D., Horowitz, A. T., Tzemach, D., Tarshish, M., Zalipsky, S., and Gabizon, A., Nuclear<br />

delivery of doxorubicin via folate-targeted liposomes with <strong>by</strong>pass of multidrug-resistantce efflux<br />

pump, Clin. Cancer Res., 6, 1949, 2000.<br />

50. Shiah, J. G., Dvorak, M., Kopeckova, P., Sun, Y., Peterson, C. M., and Kopecek, J., Biodistribution<br />

and antitumour efficacy of long-circulating N-(2-hydroxypropyl) methacrylamide copolymer-doxorubicin<br />

conjugates in nude mice, Eur. J. Cancer, 37, 131, 2001.<br />

51. Lee, E. S., Na, K., and Bae, Y. H., Doxorubicin loaded pH-sensitive polymeric micelles for reversal of<br />

resistant MCF-7 tumor, J. Control Release, 103, 405, 2005.<br />

52. Benns, J. M., Choi, J. S., Mahato, R. I., Park, J. S., and Kim, S. W., pH-sensitive cationic polymer gene<br />

delivery vehicle: N-Ac-poly(L-histidine)-graft-poly(L-lysine) comb shaped polymer, Bioconjug.<br />

Chem., 11, 637, 2000.<br />

53. Na, K. and Bae, Y. H., Self-assembled hydrogel nanoparticles responsive to tumor extracellular pH<br />

from pullulan derivative/sulfonamide conjugate: Characterization, aggregation and adriamycin<br />

release in vitro, Pharm. Res., 19, 681, 2002.<br />

54. Gottesman, M. M. and Pastan, I., Biochemistry of multidrug resistance mediated <strong>by</strong> the multidrug<br />

transporter, Ann. Rev. Biochem., 62, 385, 1993.<br />

55. Thiebaut, F., Tsuruo, T., Hamada, H., Gottesman, M. M., Pastan, I., and Willingham, M. C., Cellular<br />

localization of the multidrug-resistance gene product P-glycoprotein in normal human tissues, Proc.<br />

Natl Acad. Sci. USA, 84, 7735, 1987.<br />

56. Baldini, N., Scotlandi, K., Serra, M., Shikita, T., Zini, N., Ognibene, A., Santi, S., Ferrachini, R., and<br />

Maraldi, N. M., Nuclear immunolocalization of P-glycoprotein in multidrug-resistant cell lines<br />

showing similar mechanisms of doxorubicin distribution, Eur. J. Cell Biol., 68, 226, 1995.<br />

57. Gottesman, M. M., Pastan, I., and Ambudkar, S. V., P-glycoprotein and multidrug resistance, Curr.<br />

Opin. Genet. Dev., 6, 610, 1996.<br />

58. Mechetner, E., Kyshtoobayeva, A., Zonis, S., Kim, H., Stroup, R., Garcia, R., Parker, R. J., and<br />

Fruehauf, J. P., Levels of multidrug resistance (MDR1) P-glycoprotein expression <strong>by</strong> human<br />

breast cancer correlate with in vitro resistance to taxol and doxorubicin, Clin. Cancer Res., 4,<br />

389, 1998.<br />

59. Mahoney, B. P., Raghunand, N., Baggett, B., and Gillies, R. J., Tumor acidity, ion trapping and<br />

chemotherapeutics. I. Acid pH affects the distribution of chemotherapeutic agents in vitro, Biochem.<br />

Pharmacol., 66, 1207, 2003.<br />

60. Murthy, N., Robichaud, J. R., Tirrell, D. A., Stayton, P. S., and Hoffman, A. S., The design and<br />

synthesis of polymers for eukaryotic membrane disruption, J. Control Release, 61, 137, 1999.<br />

61. van Sluis, R., Bhujwalla, Z. M., Raghunand, N., Ballesteros, P., Alvarez, J., Cerdán, S., Galons, J. P.,<br />

and Gillies, R. J., In vivo imaging of extracellular pH using 1 H MRSI, Magn. Reso. Med., 41, 743,<br />

1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


464<br />

Nanotechnology for Cancer Therapy<br />

62. Goren, D., Horowitz, A. T., Tzemach, D., Tarshish, M., Zalipsky, S., and Gabizon, A., Nuclear<br />

delivery of doxorubicin via folate-targeted liposomes with <strong>by</strong>pass of multidrug-resistantce efflux<br />

pump, Clin. Cancer Res., 6, 1949, 2000.<br />

63. Putnam, D., Gentry, C. A., Pack, D. W., and Langer, R., Polymer-based gene delivery with low<br />

cytotoxicity <strong>by</strong> a unique balance of side-chain termini, Proc. Natl Acad. Sci., 98, 1200, 2001.<br />

64. Sabharanjak, S. and Mayor, S., Folate receptor endocytosis and trafficking, Adv. Drug Deliv. Rev., 56,<br />

1099, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


23<br />

CONTENTS<br />

cRGD-Encoded, MRI-Visible<br />

Polymeric Micelles for<br />

Tumor-Targeted Drug Delivery<br />

Jinming Gao, Norased Nasongkla, and<br />

Chalermchai Khemtong<br />

23.1 Introduction ....................................................................................................................... 465<br />

23.2 Integrin avb3-Mediated Drug Targeting to Angiogenic Tumor Vasculature .................. 466<br />

23.3 cRGD-Encoded Micelles Facilitate Doxorubicin Delivery to<br />

Tumor Endothelial Cells ................................................................................................... 467<br />

23.4 cRGD-Encoded Micelles Increase Doxorubicin Toxicity over<br />

Non-cRGD Micelles .......................................................................................................... 468<br />

23.5 Non-Invasive Imaging Methods for Pharmacokinetic Studies......................................... 469<br />

23.6 SPIO as MR T2 Contrast Agent ....................................................................................... 470<br />

23.7 Development of MRI-Visible Micelles ............................................................................ 470<br />

23.8 Non-Invasive Monitoring of Cell and Tumor Targeting <strong>by</strong><br />

cRGD-Encoded, SPIO-Loaded Micelles .......................................................................... 472<br />

23.9 Conclusions ....................................................................................................................... 473<br />

Acknowledgments ......................................................................................................................... 474<br />

References ..................................................................................................................................... 474<br />

23.1 INTRODUCTION<br />

Remarkable progress has been made in recent years to establish polymer micelles as a novel<br />

multifunctional platform of nanoscale constructs for diagnostic and therapeutic applications. 1–5<br />

Polymer micelles are composed of amphiphilic block copolymers that contain distinguished hydrophobic<br />

and hydrophilic segments. The distinct chemical nature of the two blocks results in<br />

thermodynamic phase separation in aqueous solution and formation of nanoscopic supramolecular<br />

core/shell structures. During the micellization process, the hydrophobic blocks associate to form the<br />

core region, whereas the hydrophilic blocks form the shell that separates the core from the aqueous<br />

medium. This unique architecture enables the micelle core to serve as a nanoscopic depot for<br />

therapeutic/contrast agents and the shell as biospecific surfaces for targeting applications<br />

(Figure 23.1).<br />

Biocompatible and biodegradable block copolymers are currently used in micelle design. For<br />

the hydrophilic segment, poly(ethylene glycol) (PEG) blocks with a molecular weight of 2–15 kD<br />

are most often used. Meanwhile, the hydrophobic blocks include a variety of polymers such as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

465


466<br />

Hydrophobic<br />

core<br />

Hydrated<br />

shell<br />

∅=20<br />

–200 nm<br />

FIGURE 23.1 Micellar nanomedicine platform for targeted drug delivery.<br />

Nanotechnology for Cancer Therapy<br />

Targeting ligand<br />

Iron oxide<br />

Amphiphilic polymer:<br />

Hydrophoilic block<br />

Hydrophobic block<br />

poly(D, L-lactide), poly(3-caprolactone), poly(b-benzoyl-L-aspartate), and poly(g-benzyl-Lglutamate).<br />

1,5 Different hydrophobic blocks can be used to provide different hydrophobic environments<br />

(e.g., aromatic vs. aliphatic) to maximize interactions with encapsulated agents.<br />

Pioneering work from several leading groups has significantly advanced polymeric micelles as<br />

versatile carrier systems for the delivery of hydrophobic drugs. For instance, Kataoka’s group has<br />

recently developed pH-sensitive micelles <strong>by</strong> attaching hydrophobic doxorubicin (DOX) to the<br />

polymer backbone via a hydrazone linkage. This acid-labile linkage undergoes cleavage at endosomal/lysosomal<br />

pH to release the conjugated DOX that leads to significantly improved anti-tumor<br />

efficacy. 6,7 A powerful means to improve the delivery efficiency of micelles is to introduce<br />

targeting ligands on the micelle surface to achieve an active targeting function. Among the yet<br />

limited literature reports dedicated to the active targeting of polymer-related micelle systems,<br />

recent work <strong>by</strong> Torchilin’s group provides a landmark study in this direction. 8 Torchilin and<br />

coworkers have introduced two kinds of antibodies, mAb 2C5 and mAb 2G4, to the distal PEG<br />

end of phosphatidylethanolamine–poly(ethylene glycol) (PEG–PE) micelles. Not only can the<br />

immunomicelles with cancer-specific 2C5 antibody specifically bind to cancer cells in vitro,<br />

they can also increase drug accumulation in Lewis lung carcinoma tumors in mice. When compared<br />

to administration with either taxol or non-targeting micelles, immunomicelles demonstrate the highest<br />

efficacy of tumor growth inhibition. These advances bring out the unprecedented promise that polymeric<br />

micelles can serve as effective carriers for a large variety of poorly soluble drugs.<br />

23.2 INTEGRIN avb3-MEDIATED DRUG TARGETING TO ANGIOGENIC<br />

TUMOR VASCULATURE<br />

Since the original report <strong>by</strong> Cheresh and coworkers, 9 numerous studies have demonstrated that avb3<br />

integrin is an important receptor affecting growth, local invasiveness, and metastatic potential of<br />

tumors. 10–12 In general, integrins are a family of heterodimeric membrane receptors that bind to<br />

cell-surface adhesion molecules and extracellular matrix proteins (ECM), and they play a critical<br />

role in tissue morphogenesis, repair, and remodeling. The function of integrins during angiogenesis<br />

has been most extensively studied with a vb 3 that is not readily detectable in quiescent vessels but<br />

becomes highly expressed in angiogenic vessels. 9 Recently, the crystal structures of the extracellular<br />

domains of avb3 and avb3/c(-RGDfV-) complex have been determined <strong>by</strong> Xiong et al. 13,14<br />

These structures provide molecular insights and better understanding of avb3 receptor-ligand<br />

interactions and establish the structural basis for future development of more potent and specific<br />

avb3-binding ligands. 15<br />

Vascular targeting via avb3-dependent mechanisms offers significant opportunities for the<br />

development of cancer-targeted therapeutics and diagnostics. For therapeutic applications,<br />

Ruoslahti and coworkers have conjugated DOX to a cyclic RGD peptide (RGD-4C) as identified<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


MRI-Visible Polymeric Micelles 467<br />

<strong>by</strong> a phage library. 16 The DOX-RGD-4C conjugates have demonstrated significantly improved<br />

anti-tumor efficacy in nude mice bearing human breast tumor xenografts. 17 Although quantitative<br />

pharmacokinetics of DOX-RGD-4C was not reported, it is expected that the renal clearance of the<br />

peptide-drug conjugate may be fast as a result of its small molecular weight. Recently, Ghanderhari<br />

and coworkers have shown that RGD-4C-functionalized hydroxypropylmethacrylamide (HPMA)<br />

polymer led to significantly increased tumor targeting and localization in severe combined<br />

immunodeficiency disease mice bearing DU145 or PC-3 prostate tumor xenografts over the<br />

RGE-4C-HPMA control. 18,19 This work opens up many exciting opportunities to broaden<br />

HPMA scaffold in targeted drug delivery applications. Recently, Cheresh and coworkers have<br />

reported the success of a vb 3-targeted gene therapy of cancer with cationic polymerized lipidbased<br />

nanoparticles. A mutant raf gene was delivered to tumor endothelial cells that led to<br />

tumor cell apoptosis and sustained regression of primary and metastatic melanomas in mice. 20<br />

These data provide useful precedence for introducing avb3-targeted ligands on the nano delivery<br />

systems to enhance their targeting efficiency to tumor vasculature.<br />

23.3 CRGD-ENCODED MICELLES FACILITATE DOXORUBICIN DELIVERY<br />

TO TUMOR ENDOTHELIAL CELLS<br />

Previous work in this lab has demonstrated the feasibility of producing a cyclic pentapeptide c(Arg-<br />

Gly-Asp-d-Phe-Lys, cRGDFK—also referred to as cRGD herein)-encoded polymeric micelles and<br />

micelle targeting to avb3-overexpresing tumor endothelial cells (SLK cells). 21 Figure 23.2<br />

illustrates the schematic diagram of synthesis of maleimide–terminated poly(ethylene glycol)–<br />

poly(3-caprolactone) (MAL–PEG–PCL) copolymer and cRGD attachment on the micelle<br />

surface. Doxorubicin was also encapsulated inside the micelle core. The intrinsic fluorescent<br />

properties of doxorubicin (lexZ485 nm, lemZ595 nm) permit the study of cell uptake <strong>by</strong> flow<br />

cytometry and confocal laser scanning microscopy.<br />

Figure 23.3a shows that the percentage of cell uptake increased with increasing cRGD density<br />

on the micelle surface. With 5% cRGD surface density, a modest 3-fold increase of cell uptake was<br />

O<br />

O<br />

O HO-PEG-N<br />

O<br />

ε-caprolactone HO-PEG-MAL<br />

Δ<br />

Sn(Oct) 2<br />

O H<br />

H H<br />

N<br />

HO<br />

HN<br />

O<br />

HN<br />

H<br />

O<br />

O<br />

N<br />

O<br />

cRGD-SH<br />

O<br />

O<br />

H-[OCH2CH2CH2CH2CH2C] m-(O-CH2CH2 ) n-N PCL-b-PEG-MAL<br />

SPIO<br />

Water and sonication<br />

cRGD functionalized SPIO-loaded micelle Maleimide containing micelle<br />

O<br />

NH<br />

SH<br />

O<br />

NH2 NH<br />

NH<br />

FIGURE 23.2 Synthesis of MAL–PEG–PCL copolymer and preparation of cRGD-encoded, DOX-loaded<br />

micelles. (From Nasongkla, N., Shuai, X., Ai, H., Weinberg, B. D., Pink, J., Boothman, D. A., and Gao, J.,<br />

Angew. Chem. Int. Ed. Engl., 43, 6323, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O


468<br />

Percentage of cell uptake<br />

(a)<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0 5 16 76 76+<br />

Micelle composition (% cRGD) free RGD<br />

(b) 20 μm (c)<br />

20 μm<br />

FIGURE 23.3 (See color insert following page 522.) (a) Percentage of micelle uptake in SLK tumor<br />

endothelial cells <strong>by</strong> flow cytometry as a function of cRGD density (0–76%) on the micelle surface. The last<br />

panel shows that the cell uptake of 76% cRGD-micelles is inhibited <strong>by</strong> the presence of free RGD ligands<br />

(9 mM) in solution. (b, c) Confocal laser scanning microscopy images of SLK cells treated with 0% (b) and<br />

16% (c) cRGD-micelles after incubation for 2 h. The scale bars are 20 mm in both images. (From Nasongkla,<br />

N., Shuai, X., Ai, H., Weinberg, B. D., Pink, J., Boothman, D. A., and Gao, J., Angew. Chem. Int. Ed. Engl., 43,<br />

6323, 2004. With permission.)<br />

observed. A 30-fold increase was observed <strong>by</strong> flow cytometry with 76% cRGD attachment. In the<br />

presence of excess free RGD ligands, the avb3-mediated cell uptake can be completely inhibited<br />

(last panel, Figure 23.3a). Confocal laser scanning microscopy further supports the cRGDenhanced<br />

cell uptake (Figure 23.3b and c) and demonstrates that a majority of the micelles were<br />

taken up via a vb 3-mediated endocytosis and localized mainly in the cytoplasmic compartments in<br />

the cell (as shown <strong>by</strong> the red punctated dots). Cell nuclei were stained blue <strong>by</strong> Hoechst 33342<br />

(lexZ352 nm, lemZ455 nm) and overlaid with doxorubicin fluorescent images (lexZ485 nm,<br />

lemZ595 nm).<br />

23.4 CRGD-ENCODED MICELLES INCREASE DOXORUBICIN<br />

TOXICITY OVER NON-CRGD MICELLES<br />

Nanotechnology for Cancer Therapy<br />

Figure 23.4 shows the growth inhibition of SLK cells <strong>by</strong> different micelle formulations. DOX<br />

concentration was kept the same at 1 mM for all samples. In these experiments, different DOX<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


MRI-Visible Polymeric Micelles 469<br />

Percentage of growth inhibition<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

Free DOX 0% RGD 5% RGD 16% RGD 16% RGD+<br />

16% RGD<br />

RGD w/o DOX<br />

Micelle formulation<br />

FIGURE 23.4 Cell growth in SLK cells for different compositions of SPIO-loaded micelles at 1 mM DOX<br />

concentration. The percentage of growth was calculated as the ratio of cell numbers for treated samples over<br />

untreated control. Error bars were obtained from triplicate samples.<br />

samples were exposed to SLK cells for 1 h. After 1 h, the DOX-containing media were replaced<br />

with cell culture media and kept for an additional four days until the cells in the untreated control<br />

group reached confluence. Percentage of growth (PG) was calculated as the ratio of the number of<br />

cancer cells of the treated group over the untreated control. At 1 mM DOX dose, free drug demonstrated<br />

the lowest growth at 24.4G2%. For micelle-delivered DOX, growth decreased with an<br />

increase of cRGD density on the micelle surface. The PG values for 0%, 5%, and 16% cRGDmicelles<br />

were 92G8%, 85G10%, and 51G2.58%, respectively. The cRGD micelle toxicity was<br />

inhibited <strong>by</strong> the free RGD ligand co-incubated with the micelles (PG 78G10%, Figure 23.4). In<br />

addition, cRGD-encoded, DOX-free micelles did not show much tumor growth inhibition (115G<br />

11%) at the same micelle dose. These data demonstrate the success of cRGD-encoded micelles in<br />

enhancing the drug targeting efficiency and cytotoxic response in avb3-overexpresing SLK cells.<br />

23.5 NON-INVASIVE IMAGING METHODS FOR PHARMACOKINETIC<br />

STUDIES<br />

Conventional clinical pharmacokinetic studies require the collection of blood or tissue samples to<br />

assess the clearance rate, tissue distribution, and concentration vs. time relationships of<br />

chemotherapy. Such procedures are invasive and not compliant with patients especially when<br />

multiple samples need to be collected at different time points. In addition, the quantitative interpretation<br />

of these data is complicated <strong>by</strong> individual variability and uncertainty in the technical<br />

procedure (e.g., extraction of drugs from tissue may not be complete).<br />

To circumvent these limitations, significant research efforts have focused on the development<br />

of non-invasive imaging techniques to characterize the drug pharmacokinetics in the target tissue<br />

in vivo. 22–25 Such imaging techniques include single photon emission computed tomography<br />

(SPECT), 26,27 positron-emission tomography (PET, dual photons), 26,27 and magnetic resonance<br />

imaging (MRI). 28 SPECT imaging requires the use of radionuclides (e.g., 111 In, 99m Tc) that emit<br />

high-energy gamma photons (typically 60–600 keV) whereas PET detects positron annihilation<br />

radiation (511 keV). Although these two methods provide the most sensitive measurement of<br />

radiolabeled drugs (10 K10 –10 K12 mol), they have very low spatial resolutions. For example, the<br />

resolutions of small animal SPECT and PET instruments are approximately 2 and 5 mm, respectively.<br />

26,27 Such resolution is not adequate to accurately assess drug distribution within tumor tissue,<br />

especially in a small animal tumor model (e.g., tumor xenograft in mice).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


470<br />

MRI is another imaging modality that can be used to study in situ pharmacokinetics. 28 MRI<br />

detects the magnetization changes of nuclei that have net angular momentum such as 1 H, 13 C, and<br />

19 F. Examples of drugs studied <strong>by</strong> MRI include 5-fluorouracil ( 19 F), 13 C-labelled temozolomide,<br />

and ifosfamide ( 31 P). MRI has much higher resolution (w100 mm) than nuclear imaging methods.<br />

Its main limitation is low sensitivity, making it difficult to detect drugs at their therapeutic concentrations.<br />

28 Use of exogenous MR contrast agents (e.g., SPIO) can dramatically increase the<br />

sensitivity of MR detection.<br />

23.6 SPIO AS MR T2 CONTRAST AGENT<br />

Superparamagnetic iron oxide (SPIO) nanoparticles have been extensively studied in the past<br />

decade for their use as MR contrast agents. Unlike the low molecular weight, paramagnetic<br />

metal chelates such as Gd-DTPA (T1 contrast agent) and SPIO nanoparticles are considered T2negative<br />

contrast agents with substantially higher T2 and T1 relaxivity compared to T1 agents. 29<br />

SPIO agents are composed of iron oxide nanocrystals with the general formula Fe 3C<br />

2 O 3M 2C O,<br />

where M 2C is a divalent metal ion such as iron, manganese, nickel, cobalt, or magnesium.<br />

When M 2C is ferrous iron (Fe 2C ), SPIO becomes magnetite. In the absence of an external magnetic<br />

field, the magnetic domains inside SPIO are randomly oriented with no net magnetic field. An<br />

external magnetic field can cause the magnetic dipoles of the magnetic domains to reorient, leading<br />

to dramatically increased magnetic moments and significantly shortened relaxations in both T1 and<br />

T2/T2* relaxation processes. 29<br />

SPIO nanoparticles have been classified into four different categories based on their sizes and<br />

applications: large, standard, ultrasmall, and monocrystalline agents. Large SPIO agents (O200 nm)<br />

are mainly used for oral gastrointestinal lumen imaging. For example, AMI-121 (Advanced<br />

Magnetics) is a clinically approved T2 agent with a hydrodynamic diameter of 300 nm and<br />

T2andT1relaxivitiesof72and3.2FemM K1 s K1 , respectively. 30 Standard SPIO agents<br />

(50–200 nm) are easily sequestered <strong>by</strong> the RES in the liver and spleen upon intravenous injection<br />

and are primarily used for liver and spleen imaging. The iron oxide crystal (4.8–5.6 nm) in AMI-25<br />

(Endorem w <strong>by</strong> Guerbet and Feridex w <strong>by</strong> Berlex Lab), a clinically approved agent, is coated with<br />

dextran to yield a final hydrodynamic diameter between 80 and 150 nm. The T2 and T1 relaxivities<br />

are 98.3 and 23.9 Fe mM K1 s K1 , respectively. 31 AMI-25 efficiently accumulates in the liver (w80%<br />

of injected dose) and spleen (5–10%) within minutes of administration, and it has a short blood<br />

half-life (6 min).<br />

Weissleder and coworkers have pioneered the development of ultrasmall (USPIO) and monocrystalline<br />

iron oxide nanoparticles (MION). 32–35 Two USPIO agents, AMI-227 and CLARISCAN,<br />

are currently in phase III clinical trials. AMI-227 is composed of a crystalline core of 4–6 nm<br />

surrounded <strong>by</strong> a dextran coating. Its hydrodynamic diameter is 20–40 nm, and it has T2 and T1<br />

relaxivities of 44.1 and 21.6 Fe mM K1 s K1 , respectively. 36 Because of its smaller size, AMI-227<br />

showed an increased blood half-life over 24 h. 37 Consequently, AMI-227 can be used as a blood<br />

pool agent for MR angiography. Eventually, these particles are taken up <strong>by</strong> the RES of lymph nodes<br />

and become particularly useful for MR lymphography. 38,39<br />

23.7 DEVELOPMENT OF MRI-VISIBLE MICELLES<br />

Nanotechnology for Cancer Therapy<br />

Recently, a new class of MRI-visible nanoparticles was developed through the incorporation of<br />

hydrophobic SPIO nanoparticles inside polymer micelles. 40 High loading of SPIO was achieved<br />

inside the micelle core that led to a significant increase in T2 relaxivity. The resulting SPIO-loaded<br />

micelles have shown superb sensitivity under a clinical 1.5 T MR scanner.<br />

First, monodisperse SPIO (Fe 3O 4) nanoparticles with different sizes were synthesized<br />

following published work <strong>by</strong> Sun and coworkers. 41 Transmission electron microscopy (TEM)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


MRI-Visible Polymeric Micelles 471<br />

(a)<br />

(b)<br />

(c)<br />

Volume<br />

Volume<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

(d)<br />

Volume<br />

0<br />

11<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

11<br />

(e)<br />

(f)<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

11<br />

17 26 41 62 89 119<br />

Diameter (nm)<br />

17 26 41 60 80 108<br />

Diameter (nm)<br />

17 26 41 60 80 110<br />

Diameter (nm)<br />

FIGURE 23.5 TEM images and DLS data of different SPIO-micelle formulations. (a, d) Single SPIO (4 nm in<br />

diameter) encapsulated <strong>by</strong> PEG5k–DSPE lipid; (b, e) SPIO (4 nm)-loaded PEG5k-b-PCL5k micelles; (c, f)<br />

SPIO (8 nm)-loaded PCL5k-b-PEG5k micelles. A cluster of SPIO particles are loaded inside each PCL5k-b-<br />

PEG5k micelle as shown in (b) and (c). (From Ai, H., Flask, C., Weinberg, B., Shuai, X., Pagel, M., Farrell, D.,<br />

Duerk, J., and Gao, J., Adv. Mat., 17, 1949, 2005. With permission.)<br />

characterization of SPIO nanoparticles showed 4, 8, and 16 nm in diameter. The size distribution is<br />

uniform for all particle sizes. Distearoylphosphatidyl–ethanolamine–poly(ethylene glycol)<br />

(PEG5k–DSPE) lipid and amphiphilic diblock copolymer of PCL5k-b-PEG5k were used for the<br />

micelle formation. Interestingly, PEG5k–DSPE lipid micelles provide individual encapsulation of<br />

single 4 nm SPIO nanoparticles (Figure 23.5a). In comparison, PEG5k-b-PCL5k polymer micelles<br />

allow the incorporation of a cluster of 4 and 8 nm SPIO nanoparticles, respectively (Figure 23.5b<br />

and Figure 23.5c). Dynamic light scattering (DLS) was used to study the size distribution of SPIOloaded<br />

PEG5k-b-PCL5k polymer micelles. DLS experiments were performed using the micelle<br />

solutions with 908 scattering angle at 258C. Figure 23.5d shows a representative DLS diagram of<br />

PEG5k–DSPE micelles containing 4 nm SPIO nanoparticles. The mean hydrodynamic diameter<br />

was measured to be 17G1 nm. The mean hydrodynamic diameters were 75G4 and 97G6 nm for<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


472<br />

4 nm SPIO<br />

DSPE-PEG5k<br />

4 nm SPIO<br />

PCL5k-PEG5k<br />

Fe Concentration (μM)<br />

1040 520 260 130 65 32.5 16.25 H 2O<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 23.6 T2-weighted MRI scan (TRZ5000 ms, TEZ40 ms) of single 4 nm SPIO encapsulated <strong>by</strong><br />

PEG5k-DSPE and a cluster of 4 nm SPIO loaded in PEG5k-PCL5k micelles. MRI intensity at the same Fe<br />

concentration is compared. (From Ai, H., Flask, C., Weinberg, B., Shuai, X., Pagel, M., Farrell, D., Duerk, J.,<br />

and Gao, J., Adv. Mat., 17, 1949, 2005. With permission.)<br />

PEG-b-PCL micelles containing 4 and 8 nm SPIO particles (Figure 23.5e and Figure 23.5f),<br />

respectively.<br />

The T1 and T2 relaxivities of different micelle particles were determined on a clinical 1.5 T<br />

scanner (Siemens Sonata). The T1 relaxivities for different micelle formulations were comparable<br />

and ranged from 1.3 to 2.9 Fe mM K1 s K1 . In contrast, the T2 relaxivities of micelles notably<br />

depended on SPIO clustering, SPIO diameter, and loading density. Figure 23.6 illustrates the<br />

MR signal intensity as a function of Fe concentration for two micelle formulations. The top row<br />

shows SPIO (4 nm) particles individually encapsulated <strong>by</strong> PEG5k–DSPE lipid (sample in<br />

Figure 23.5a), and the bottom row shows PEG-b-PCL micelles loaded with the same 4 nm SPIO<br />

particles (sample in Figure 23.5b). At the same Fe concentration, SPIO-loaded PEG-b-PCL<br />

micelles showed significantly increased contrast in these T2-weighted images. The T2 relaxivity<br />

increased from 25 for single SPIO micelles to 169 Fe mM K1 s K1 for SPIO-loaded polymeric<br />

micelles (w7 fold increase). Clustering of SPIO particles considerably increased the T2 relaxivity<br />

per Fe concentration. For polymeric micelles, further increase in SPIO diameter (e.g., from 4 to 8<br />

and 16 nm) also led to significantly increased loading density and enhanced T2 relaxivity. The<br />

values of T2 relaxivity reached 318 and 471 Fe mM K1 s K1 for micelles loaded with 8 and 16 nm<br />

Fe3O4 particles, respectively. 40<br />

The MR sensitivity of detection was measured for the micelle formulations using a T2weighted<br />

scan (TRZ5000 ms, TEZ40 ms) in an in vitro test tube experiment. The sensitivity<br />

limit is defined as the micelle concentration where the MR signal intensity decreases to 50% of the<br />

signal of pure water. For PEG-b-PCL micelles loaded with 16 nm SPIO particles, the highest<br />

sensitivity limit was observed at 5.2 mg/mL. In this micelle formulation, the micelle diameter is<br />

110G9 nm, and Fe3O4 loading is 54.2%. Because the molecular weight of typical micelles is on the<br />

order of 10 6 g/mol, 2 the above sensitivity limit corresponds to a micelle concentration of approximately<br />

5 nM.<br />

23.8 NON-INVASIVE MONITORING OF CELL AND TUMOR TARGETING<br />

BY CRGD-ENCODED, SPIO-LOADED MICELLES<br />

Recently, cRGD-encoded, SPIO-loaded polymer micelles using similar procedures were successfully<br />

developed as shown in Figure 23.2. In this system, 8 nm SPIO nanoparticles were<br />

encapsulated (6.7 wt.%) into cRGD-encoded (50%) and non-cRGD (0%) PEG–PLA micelles.<br />

The mean hydrodynamic diameters of both micelle formulations are approximately 40 nm <strong>by</strong><br />

dynamic light scattering measurement. In cell uptake experiments, tumor SLK endothelial cells<br />

(1.2!10 6 cells) were co-incubated with SPIO micelles at the final iron concentrations of 0, and<br />

12.5 mg/mL in 10% fetal bovine serum cell culture medium for two hours. The medium was then<br />

removed, and the cells were gently washed twice with HBSS. Cells were detached <strong>by</strong> trypsin, and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


MRI-Visible Polymeric Micelles 473<br />

MRI Intensity<br />

(a)<br />

15<br />

10<br />

cell pellets were collected after centrifugation at 1200 rpm for 5 min. Cells were then fixed<br />

with 2% paraformaldehyde in phosphate-buffered saline (PBS) (200 mL) for two hours and<br />

concentrated <strong>by</strong> centrifugation again. Finally, each sample was dispersed in 20 mL of 2%<br />

agarose gel in PBS. A 384-well plate was used to contain the samples for MRI imaging. MR<br />

images were collected with conventional T2-weighted spin echo acquisition parameters (TRZ<br />

6,000 ms, TEZ90 ms).<br />

Figure 23.7 shows the MR signal intensity of SLK cells incubated with cRGD-encoded and<br />

non-cRGD micelles. Three major conclusions can be drawn from these experiments. First, cRGDencoded<br />

micelles led to significant reduction of MR signal amplitude over non-cRGD micelles as a<br />

result of the increased micelle uptake in SLK cells. Second, non-specific uptake of non-cRGD<br />

micelles is relatively small with minimal change in MR intensity over the Fe concentration range<br />

from 0 to 12.5 mg/mL. Third, MR detection is highly sensitive in imaging tumor cells (1.2!10 6<br />

SLK cells at 12!10 6 cells/mL) with specific avb3-mediated uptake of Fe3O4 micelles. At 12.5<br />

Fe mg/mL, MR amplitude decreased from 9.44 for non-cRGD micelles to 4.98 for cRGD-encoded<br />

micelles. The above data demonstrate the feasibility in producing biologically specific, highly<br />

sensitive MR molecular probes based on the micelle construct and design.<br />

23.9 CONCLUSIONS<br />

5<br />

0<br />

50 %<br />

0 %<br />

12.5 mg/mL SPIO Cell PBS<br />

FIGURE 23.7 (a) MRI signal intensity of SLK cells treated with SPIO-loaded PEG–PLA micelles. SLK cells<br />

with micelle treatment and PBS buffer were used as control. Images (b) and (c) correspond to the cells<br />

incubated with 50% and 0% cRGD-micelles at 12.5 mg Fe/mL, respectively. (d) MR image of the vial<br />

containing SLK cells without any micelle incubation. (e) MR image of the vial containing PBS. MRI intensity<br />

is significantly darkened in cells incubated with cRGD-encoded micelles (b).<br />

The long-term goal of this research is to establish polymer micelles as a safe and efficacious<br />

nanomedicine platform for targeted cancer therapy. The cRGD-encoded micelles have shown<br />

superb targeting efficiency in cell culture in vitro that establishes the micelle foundation for<br />

future in vivo evaluations in tumor-bearing mice. To facilitate the evaluation of tumor targeting<br />

efficiency of these micelles, MRI-visible micelles have been developed. Clustering of SPIO<br />

particles inside the micelle core led to a dramatic increase in T2 relaxivity. Together with high<br />

loading density of Fe3O4 inside polymer micelles, they led to an ultra sensitivity of detection <strong>by</strong><br />

MRI per micelle basis (w5 nM). The much improved sensitivity of SPIO-loaded micelles may<br />

prove essential in detecting low concentrations of micelle nanoparticles for in vivo intratumoral<br />

tissue distribution studies <strong>by</strong> MRI.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

(b)<br />

(c)<br />

(d)<br />

(e)


474<br />

ACKNOWLEDGMENTS<br />

This work was supported <strong>by</strong> the National Institutes of Health grants R01 CA90696 and R21<br />

EB005394 to Jinming Gao. Norased Nasongkla acknowledges the Royal Thai Government for a<br />

predoctoral fellowship support.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Adams, M. L., Lavasanifar, A., and Kwon, G. S., Amphiphilic block copolymers for drug delivery,<br />

J. Pharm. Sci., 92, 1343–1355, 2003.<br />

2. Kwon, G. S., Polymeric micelles for delivery of poorly water-soluble compounds, Crit. Rev. Ther.<br />

Drug. Carrier Syst., 20, 357–403, 2003.<br />

3. Lukyanov, A. N. and Torchilin, V. P., Micelles from lipid derivatives of water-soluble polymers as<br />

delivery systems for poorly soluble drugs, Adv. Drug Deliv. Rev., 56, 1273–1289, 2004.<br />

4. Otsuka, H., Nagasaki, Y., and Kataoka, K., PEGylated nanoparticles for biological and pharmaceutical<br />

applications, Adv. Drug Deliv. Rev., 55, 403–419, 2003.<br />

5. Torchilin, V. P., Targeted polymeric micelles for delivery of poorly soluble drugs, Cell. Mol. Life Sci.,<br />

61, 2549–2559, 2004.<br />

6. Bae, Y., Fukushima, S., Harada, A., and Kataoka, K., Design of environment-sensitive supramolecular<br />

assemblies for intracellular drug delivery: Polymeric micelles that are responsive to intracellular<br />

pH change, Angew. Chem. Int. Ed. Engl., 42, 4640–4643, 2003.<br />

7. Bae, Y., Nishiyama, N., Fukushima, S., Koyama, H., Yasuhiro, M., and Kataoka, K., Preparation and<br />

biological characterization of polymeric micelle drug carriers with intracellular pH-triggered drug<br />

release property: Tumor permeability, controlled subcellular drug distribution, and enhanced in vivo<br />

antitumor efficacy, Bioconjug. Chem., 16, 122–130, 2005.<br />

8. Torchilin, V. P., Lukyanov, A. N., Gao, Z., and Papahadjopoulos-Sternberg, B., Immunomicelles:<br />

Targeted pharmaceutical carriers for poorly soluble drugs, Proc. Natl. Acad. Sci. USA, 100,<br />

6039–6044, 2003.<br />

9. Brooks, P. C., Clark, R. A., and Cheresh, D. A., Requirement of vascular integrin alpha v beta 3 for<br />

angiogenesis, Science, 264, 569–571, 1994.<br />

10. Ruegg, C., Dormond, O., and Foletti, A., Suppression of tumor angiogenesis through the inhibition of<br />

integrin function and signaling in endothelial cells: Which side to target? Endothelium, 9, 151–160, 2002.<br />

11. Teti, A., Migliaccio, S., and Baron, R., The role of the alphaVbeta3 integrin in the development of<br />

osteolytic bone metastases: A pharmacological target for alternative therapy? Calcif. Tissue Int., 71,<br />

293–299, 2002.<br />

12. Varner, J. A. and Cheresh, D. A., Tumor angiogenesis and the role of vascular cell integrin alphavbeta3,<br />

Important Adv. Oncol., 69–87, 1996.<br />

13. Xiong, J. P., Stehle, T., Diefenbach, B., Zhang, R., Dunker, R., Scott, D. L., Joachimiak, A.,<br />

Goodman, S. L., and Arnaout, M. A., Crystal structure of the extracellular segment of integrin<br />

alpha Vbeta3, Science, 294, 339–345, 2001.<br />

14. Xiong, J. P., Stehle, T., Zhang, R., Joachimiak, A., Frech, M., Goodman, S. L., and Arnaout, M. A.,<br />

Crystal structure of the extracellular segment of integrin alpha V beta3 in complex with an Arg–Gly–<br />

Asp ligand, Science, 296, 151–155, 2002.<br />

15. Gottschalk, K. E. and Kessler, H., The structures of integrins and integrin-ligand complexes: Implications<br />

for drug design and signal transduction, Angew. Chem. Int. Ed. Engl., 41, 3767–3774, 2002.<br />

16. Pasqualini, R., Koivunen, E., and Ruoslahti, E., Alpha v integrins as receptors for tumor targeting <strong>by</strong><br />

circulating ligands, Nat. Biotechnol., 15, 542–546, 1997.<br />

17. Arap, W., Pasqualini, R., and Ruoslahti, E., Cancer treatment <strong>by</strong> targeted drug delivery to tumor<br />

vasculature in a mouse model, Science, 279, 377–380, 1998.<br />

18. Line, B. R., Mitra, A., Nan, A., and Ghandehari, H., Targeting tumor angiogenesis: Comparison of<br />

Peptide and polymer-Peptide conjugates, J. Nucl. Med., 46, 1552–1560, 2005.<br />

19. Mitra, A., Mulholland, J., Nan, A., McNeill, E., Ghandehari, H., and Line, B. R., Targeting tumor<br />

angiogenic vasculature using polymer-RGD conjugates, J. Control. Release, 102, 191–201, 2005.<br />

20. Hood, J. D., Bednarski, M., Frausto, R., Guccione, S., Reisfeld, R. A., Xiang, R., and Cheresh, D. A.,<br />

Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science, 296, 2404–2407, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


MRI-Visible Polymeric Micelles 475<br />

21. Nasongkla, N., Shuai, X., Ai, H., Weinberg, B. D., Pink, J., Boothman, D. A., and Gao, J., cRGDfunctionalized<br />

polymer micelles for targeted doxorubicin delivery, Angew. Chem. Int. Ed. Engl., 43,<br />

6323–6327, 2004.<br />

22. Fischman, A. J., Alpert, N. M., and Rubin, R. H., Pharmacokinetic imaging: A noninvasive method for<br />

determining drug distribution and action, Clin. Pharmacokinet., 41, 581–602, 2003.<br />

23. Port, R. E. and Wolf, W., Noninvasive methods to study drug distribution, Invest. New Drugs, 21,<br />

157–168, 2003.<br />

24. Seddon, B. M. and Workman, P., The role of functional and molecular imaging in cancer drug<br />

discovery and development, Br. J. Radiol., 76, S128–S138, 2003.<br />

25. Singh, M. and Waluch, V., Physics and instrumentation for imaging in-vivo drug distribution, Adv.<br />

Drug Deliv. Rev., 41, 7–20, 2000.<br />

26. Bhatnagar, A., Hustinx, R., and Alavi, A., Nuclear imaging methods for non-invasive drug monitoring,<br />

Adv. Drug Deliv. Rev., 41, 41–54, 2000.<br />

27. Saleem, A., Aboagye, E. O., and Price, P. M., In vivo monitoring of drugs using radiotracer techniques,<br />

Adv. Drug Deliv. Rev., 41, 21–39, 2000.<br />

28. Griffiths, J. R. and Glickson, J. D., Monitoring pharmacokinetics of anticancer drugs: Non-invasive<br />

investigation using magnetic resonance spectroscopy, Adv. Drug Deliv. Rev., 41, 75–89, 2000.<br />

29. Wang, Y., Hussain, S., and Krestin, G., Superparamagnetic iron oxide contrast agents: Physicochemical<br />

characteristics and applications in MR imaging, Eur. Radiol., 11, 2319–2331, 2001.<br />

30. Hahn, P. F., Stark, D. D., Lewis, J. M., Saini, S., Elizondo, G., Weissleder, R., Fretz, C. J., and<br />

Ferrucci, J. T., First clinical trial of a new superparamagnetic iron oxide for use as an oral gastrointestinal<br />

contrast agent in MR imaging, Radiology, 175, 695–700, 1990.<br />

31. Weissleder, R., Stark, D. D., Engelstad, B. L., Bacon, B. R., Compton, C. C., White, D. L., Jacobs, P.,<br />

and Lewis, J., Superparamagnetic iron oxide: Pharmacokinetics and toxicity, AJR Am. J. Roentgenol.,<br />

152, 167–173, 1989.<br />

32. Moore, A., Weissleder, R., and Bogdanov, A. Jr., Uptake of dextran-coated monocrystalline iron<br />

oxides in tumor cells and macrophages, J. Magn. Reson. Imaging, 7, 1140–1145, 1997.<br />

33. Schaffer, B. K., Linker, C., Papisov, M., Tsai, E., Nossiff, N., Shibata, T., Bogdanov, A. Jr., Brady,<br />

T. J., and Weissleder, R., MION-ASF: Biokinetics of an MR receptor agent, Magn. Reson. Imaging,<br />

11, 411–417, 1993.<br />

34. Weissleder, R., Elizondo, G., Wittenberg, J., Rabito, C. A., Bengele, H. H., and Josephson, L.,<br />

Ultrasmall superparamagnetic iron oxide: Characterization of a new class of contrast agents for<br />

MR imaging, Radiology, 175, 489–493, 1990.<br />

35. Weissleder, R., Lee, A. S., Fischman, A. J., Reimer, P., Shen, T., Wilkinson, R., Callahan, R. J., and<br />

Brady, T. J., Polyclonal human immunoglobulin G labeled with polymeric iron oxide: Antibody MR<br />

imaging, Radiology, 181, 245–249, 1991.<br />

36. Jung, C. W. and Jacobs, P., Physical and chemical properties of superparamagnetic iron oxide MR<br />

contrast agents: Ferumoxides, ferumoxtran, ferumoxsil, Magn. Reson. Imaging, 13, 661–674, 1995.<br />

37. McLachlan, S. J., Morris, M. R., Lucas, M. A., Fisco, R. A., Eakins, M. N., Fowler, D. R., Scheetz,<br />

R. B., and Olukotun, A. Y., Phase I clinical evaluation of a new iron oxide MR contrast agent, J. Magn.<br />

Reson. Imaging, 4, 301–307, 1994.<br />

38. Anzai, Y., Blackwell, K. E., Hirschowitz, S. L., Rogers, J. W., Sato, Y., Yuh, W. T., Runge, V. M.,<br />

Morris, M. R., McLachlan, S. J., and Lufkin, R. B., Initial clinical experience with dextran-coated<br />

superparamagnetic iron oxide for detection of lymph node metastases in patients with head and neck<br />

cancer, Radiology, 192, 709–715, 1994.<br />

39. Anzai, Y., Brunberg, J. A., and Lufkin, R. B., Imaging of nodal metastases in the head and neck,<br />

J. Magn. Reson. Imaging, 7, 774–783, 1997.<br />

40. Ai, H., Flask, C., Weinberg, B., Shuai, X., Pagel, M., Farrell, D., Duerk, J., and Gao, J., Magnetiteloaded<br />

polymeric micelles as novel magnetic resonance probes, Adv. Mat., 17, 1949–1952, 2005.<br />

41. Sun, S. and Zeng, H., Size-controlled synthesis of magnetite nanoparticles, J. Am. Chem. Soc., 124,<br />

8204–8205, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


24<br />

CONTENTS<br />

Targeted Antisense<br />

Oligonucleotide Micellar<br />

Delivery Systems<br />

Ji Hoon Jeong, Sun Hwa Kim, and Tae Gwan Park<br />

24.1 Introduction ....................................................................................................................... 477<br />

24.2 Micellar Antisense Oligonucleotide Delivery Systems .................................................... 478<br />

24.2.1 Poly(D, L-Lactic-Co-Glycolic Acid)-Oligonucleotides<br />

Hybrid Polymeric Micelles ................................................................................. 478<br />

24.2.2 Oligonucleotide Delivery Systems Based on Polyelectrolyte<br />

Complex Micelles ............................................................................................... 478<br />

24.2.3 In Vivo Applications ........................................................................................... 481<br />

24.3 Ligand-Mediated Targeting of Antisense Oligonucleotides ............................................ 483<br />

24.3.1 Ligand-Mediated ODN Targeting Systems Based on PEC Micelles ................ 483<br />

24.3.2 In Vivo Applications ........................................................................................... 483<br />

24.4 Conclusions ....................................................................................................................... 484<br />

References..................................................................................................................................... 484<br />

24.1 INTRODUCTION<br />

Antisense oligonucleotides (antisense ODNs) are receiving increasing attention because of their<br />

high selectivity toward the sequence of a target mRNA. 1–4 The sequence-specific binding results in<br />

the degradation of the target mRNA, leading to blockage of the target protein expression. Owing to<br />

the features, antisense ODNs have been popularly employed for the modulation of specific gene<br />

expression. The clinical use of antisense ODN, however, is still limited as a result of its intrinsic<br />

properties: vulnerability to enzymatic digestion <strong>by</strong> humoral and cellular nucleases, and poor<br />

cellular uptake in a target tissue. 5–7 The susceptibility of ODN to enzymatic digestion has been<br />

partly improved <strong>by</strong> modifying the ODN backbone with either phosphorothioate linkage or amide<br />

linkage peptide nucleic acid (PNA). 8–10 The problem with poor cellular uptake could be addressed<br />

<strong>by</strong> employing cationic lipids and polymers. 11–15 The electrostatic interaction between a cationic<br />

lipid/polymer and negatively charged DNA or ODN can lead to the formation of polyelectrolyte<br />

complex (PEC) nanoaggregates that can greatly facilitate cellular uptake of the polyanions <strong>by</strong> an<br />

adsorption-induced internalization process called endocytosis. Despite their satisfactory performance<br />

in intracellular gene transfer in vitro, the use of PEC is still limited because of its<br />

rapid clearance and undesirable pharmacokinetic profiles upon intravenous administration. 16,17<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

477


478<br />

A series of recent developments for micellar gene delivery systems could provide a chance to<br />

overcome the potential problems concerning in vivo application of PEC systems. Self-assembling<br />

polymeric micelles have been considered as potential drug carriers for anti-cancer drugs, including<br />

nucleic acid-based therapeutic drugs such as antisense ODNs. Most polymeric micelles generally<br />

have a core-shell structure composed of hydrophobic parts as an internal core and hydrophilic parts<br />

as a surrounding corona. The hydrophilic corona could protect micelles from nonspecific adsorption<br />

of serum proteins in the blood stream, leading to an avoidance of opsonization-induced clearance of<br />

the micelles. The size of polymeric micelles, usually less than 100 nm, provides them with a chance<br />

to escape from renal exclusion and reticulo-endothelial system (RES). Because of the small size of<br />

the micelles, the enhanced permeability through loosened endothelial cell junctions in the vicinity<br />

of solid tumors can also be expected. A hybrid conjugate of ODN with hydrophobic biodegradable<br />

poly(D, L-lactic-co-glycolic acid) (PLGA) that exhibited micellar properties in aqueous solution<br />

was recently introduced. 6 PEC micelles are also a new class of gene delivery systems based on<br />

electrostatically interacted DNA/polymer complexes. The PEC micelles could be prepared from the<br />

interaction between ODN and cationic block copolymers, including poly(ethylene glycol) (PEG)-bpoly(L-lysine)<br />

di-block copolymer and PEG-grafted polyethylenimine (PEI). 18,19 Alternatively, the<br />

PEC micelles could be also formed <strong>by</strong> interacting PEG-conjugated ODN (ODN–PEG) and cationic<br />

polymers or peptides. 20 These systems could provide ODN with the enhanced solubility of nanocarriers<br />

and an improved stability against nuclease attacks. Introduction of cell binding ligands such<br />

as transferring and folate could also enhance the cellular uptake and improve the in vivo tissue<br />

distribution of the PEC micelles.<br />

24.2 MICELLAR ANTISENSE OLIGONUCLEOTIDE DELIVERY SYSTEMS<br />

24.2.1 POLY(D, L-LACTIC-CO-GLYCOLIC ACID)-OLIGONUCLEOTIDES<br />

HYBRID POLYMERIC MICELLES<br />

Nanotechnology for Cancer Therapy<br />

Micelle-forming amphiphilic copolymers have been utilized as a solubilizer and carrier of various<br />

hydrophobic drugs, including several water insoluble anti-cancer drugs. An A–B type block copolymer<br />

of PEG-b-PLGA, containing doxorubicin conjugated to the terminal end of PLGA, could<br />

form micelles and remarkably improve the cellular uptake of doxorubicin. 21 Slow degradation of<br />

PLGA chains permit sustained release of the conjugated drug over a desired period. An alternative<br />

micelle-forming copolymer composed of PLGA and ODN (ODN/PLGA) was synthesized for<br />

intracellular ODN delivery. 6 Water-insoluble PLGA segments serve as a hydrophobic core and<br />

hydrophilic ODN as a surrounding corona, resulting in an amphiphilic structure similar to A–B type<br />

di-block copolymers. The ODN/PLGA could form spherical micelles with a size of ca. 80 nm. The<br />

micelles have a critical micelle concentration (cmc) ranging from 5 to 10 mg/L, depending on the<br />

analytical methods. When ODN/PLGA micelles were subjected to incubation in an aqueous<br />

solution, the micelles released ODN in a sustained manner <strong>by</strong> controlled degradation of conjugated<br />

PLGA chains. The micelles demonstrated much higher intracellular delivery efficiency in NIH3T3<br />

mouse fibroblast cells when compared to naked ODN. The fluorescently labeled micelles were<br />

distributed over the cytoplasm of the cells, suggesting the micelles were taken up <strong>by</strong> the cells<br />

via endocytosis.<br />

24.2.2 OLIGONUCLEOTIDE DELIVERY SYSTEMS BASED ON POLYELECTROLYTE COMPLEX MICELLES<br />

Electrostatic interaction between oppositely charged polyelectrolytes (e.g., polycations and DNA)<br />

leads to the formation of nanoparticular polyelectrolyte complex (PEC) <strong>by</strong> charge neutralization.<br />

The nanoparticulates are usually unstable in aqueous milieu as a result of the high incidence of<br />

collision among particles and their high surface energy. Because they tend to aggregate one<br />

another, often resulting in precipitation, the complex formation between polycations and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Targeted Antisense Oligonucleotide Micellar Delivery Systems 479<br />

polyanions (ODN or plasmid DNA) should be nonstoichiometric; there the surface charge of the<br />

complexes should be negative (polyanions excess) or positive (polycation excess). The nonstoichiometric<br />

condition between polycation and polyanion can prevent the premature inter-particular<br />

aggregation <strong>by</strong> charge repulsion among the complexes during their formation. 22 However, the<br />

complexes with negatively charged surfaces are not appropriate for intracellular gene transfer<br />

because of the negatively charged characteristics of the cellular membrane. The positively<br />

charged particles are also undesirable for in vivo application because of their susceptibility to<br />

the nonspecific interaction of serum components such as negatively charged albumin that could<br />

often lead to inter-particular aggregation in the blood stream or opsonization-mediated clearance<br />

<strong>by</strong> reticuloendothelial cells. The introduction of hydrophilic PEG segments to the complexes’<br />

surfaces greatly increases their stability, not only enhancing the solubility of the complexes,<br />

but also efficiently preventing inter-particular aggregation <strong>by</strong> the steric effect of the flexible PEG<br />

chains. 23,24 Depending on the length (molecular weight) and density (the number of PEG present on<br />

the surface), the PEG chains around the PEC are often enough to shield the surface charge of the<br />

nonstoichiometric complexes to achieve charge neutrality of the complex surface. 25–28<br />

Block or graft copolymers consisting of a cationic polymer and PEG were reported to form PEC<br />

micelles <strong>by</strong> interacting with negatively charged ODN or plasmid DNA. When a poly(1-lysine)-PEG<br />

(PLL-b-PEG) di-block copolymer interacts with negatively charged ODN or plasmid DNA, selfassembled<br />

micelles can generate. 18 The charge neutralization between polycation and negatively<br />

charged polynucleotide could form a core-shell type nanoparticulate, entrapped and surrounded <strong>by</strong><br />

a hydrophilic PEG segment. Similarly, the formation of self-associated micelles based on the<br />

interaction between counter ions could also be observed in a mixture of PEG-grafted PEI and<br />

alkyl sulfates. 19 The surrounded hydrophilic shell not only stabilizes the micelles, but also increase<br />

water solubility of the particles. The PEC micelles formed from PLL-b-PEG, and ODN showed a<br />

spherical shape with a diameter of about 50 nm as determined <strong>by</strong> dynamic light scattering. The<br />

formation of PEC micelles also greatly enhanced the stability of particles. Overnight incubation at<br />

room temperature and even storage under frozen conditions did not affect the size distribution of the<br />

micelles. 18 The addition of oppositely charged poly(aspartic acid) to PLL-b-PEG/DNA PEC<br />

micelles caused exchange reaction between poly(aspartic acid) and DNA, leading to a release of<br />

DNA in the medium. 29 This result suggested that the PEC micelle-based delivery system could<br />

stabilize the complex and increase its water solubility <strong>by</strong> efficiently compartmentalizing it within<br />

PEG shells while retaining its DNA releasing capability <strong>by</strong> exchange reaction with counter ions.<br />

The PEC micelles also efficiently protect the encapsulated ODN or plasmid DNA from enzymatic<br />

attack. 30,31 Environmental stimuli-sensitive PEC micelles were prepared using thiolated PLL-b-<br />

PEG. The micelles formed from thiolated PLL-b-PEG and ODN were dissociated to release ODN<br />

in the presence of reduced glutathione, suggesting that the ODN entrapped in PEC micelles could<br />

be released in a cytoplasmic environment for proper therapeutic activity of ODN. 32 Other polyamine-PEG<br />

conjugates including PEI–PEG and polyspermine (PSP)-PEG also showed PEC<br />

micelles forming capability when introduced to ODN. 33,34 Diameters of the PEC micelles from<br />

PEI–PEG and PSP–PEG were ca. 32 nm and ca. 12 nm as determined <strong>by</strong> dynamic light scattering<br />

and transmission electron microscopy (TEM). The micelles could be stably stored in a solution for<br />

several months. The micellar formulation could efficiently minimize the nonspecific interaction<br />

with a major component of serum proteins, namely, bovine serum albumin (BSA).<br />

Recently, an alternative class of PEC micelles-based ODN delivery systems where PEC<br />

micelles could form from the interaction between PEG-conjugated oligonucleotide and polycation<br />

was introduced. Figure 24.1 shows the schematic diagram of the formation of PEC micelles. An<br />

A–B block-type conjugate, ODN–PEG, was synthesized. 20 An acid labile phosphoroamidate<br />

linkage was employed between ODN and PEG as a stimuli-sensitive linkage for the efficient<br />

release of ODN from the micelles in acidic endosomal pH. The ODN–PEG conjugate could<br />

self-associate to form PEC micelles, <strong>by</strong> interacting with several functional polycations such as<br />

fusogenic KALA peptide, PEI, and protamine (Figure 24.2). The charge-neutralized<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


480<br />

O<br />

HN<br />

O<br />

O O O N O<br />

H H O<br />

H<br />

H OH<br />

ODN<br />

O<br />

HN<br />

O<br />

O O O N O<br />

H<br />

H<br />

O<br />

H OH<br />

H<br />

N<br />

O<br />

PEG<br />

Cationic polymer<br />

FIGURE 24.1 The schematic diagram of the formation of PEC micelles.<br />

Protective PEG<br />

shell<br />

Polyelectrolyte<br />

complex core<br />

ODN/polycation complex was entrapped and compartmentalized <strong>by</strong> surrounding PEG corona to<br />

form a stable core-shell structure. The spontaneously formed PEC micelles showed a spherical<br />

shape with a diameter of ca. 70 nm and a narrow distribution. 35 The PEC micellar formulation<br />

demonstrated much higher intracellular transport efficiency when compared to naked ODN. When<br />

formulated with antisense ODN corresponding to c-myc or c-raf mRNA, the PEC micelles could<br />

remarkably inhibit the proliferation of smooth muscle cells (A7R5) or ovarian cancer cells (A2780),<br />

respectively. 35,36 The formation of PEC micelles between ODN–PEG conjugate and linear PEI<br />

provides ODN with enhanced nuclease resistance. 37 The PEC formulation also showed minimal<br />

adsorption of serum proteins, suggesting that the surrounding PEG shell could provide the nanoparticulate<br />

with improved water solubility and with an efficient protection from the nonspecific or<br />

Percent cell growth (%)<br />

100<br />

80<br />

60<br />

40<br />

0<br />

KALA<br />

PEI Protamine no ODN<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 24.2 Influence of PEC micelles containing antisense c-myb on the proliferation of smooth muscle<br />

cells. Each PEC micelle was prepared <strong>by</strong> ODN–PEG conjugate with different polycations, KALA, PEI, and<br />

protamine. The size and surface zeta-potential values of the different micellar formulations were similar to<br />

those of the ODN–PEG/KALA formulation. The concentration of antisense ODN used in each formulation<br />

was 20 mg/ml. (From Jeong, J. H., Kim, S. W., and Park, T. G., Bioconjug. Chem., 14, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Targeted Antisense Oligonucleotide Micellar Delivery Systems 481<br />

specific interaction with extracellular or intracellular proteins. These features would make PEC<br />

micellar formulation feasible for in vivo applications, especially for a systemic administration.<br />

24.2.3 IN VIVO APPLICATIONS<br />

There are two important factors that should be considered in designing polymeric gene carriers for<br />

in vivo application: pharmacokinetic behaviors in systemic circulation and cellular uptake and<br />

intracellular trafficking of the carriers. Systemically administered polymeric gene carriers could<br />

experience a number of barriers upon injection, including the attacks of extracellular nucleases and<br />

systemic clearance systems such as RES. The accumulation of the carriers in a desired tissue or<br />

organ could also be one of the barriers the carriers should overcome. Even after arriving at the<br />

appropriate destination, there are more cellular barriers, including cellular uptake, escape from<br />

endosomal compartment, and localization into the nucleus when it is required. A number of<br />

previous reports on micelle-based drug delivery systems demonstrated improved pharmacokinetics<br />

and cellular internalization. 21,38–42 The PEC micelles have a similar core-shell structure to polymeric<br />

micelles. The surrounding hydrophilic shell could minimize the nonspecific adsorption of<br />

serum proteins to avoid the surveillance of the systemic clearance systems, allowing prolonged<br />

circulation of the carrier. Also, the colloidal size of PEC micelles makes it feasible for them to be<br />

used for passive tumor targeting. Nanoparticulates, generally having less than 100 nm, could be<br />

expected to experience enhanced permeability and retention in the vicinity of a solid tumor. 43 In<br />

addition, the use of functional polymers that can facilitate the endosomal escape of ODN will be<br />

advantageous to improve intracellular trafficking. For example, PEI, one of the commonly used<br />

cationic gene carriers, has a high buffering capacity that may facilitate escape of the complexes<br />

from the endosomal compartment. 44<br />

In vivo anti-tumor activity of the PEC micellar formulation (ODN–PEG/PEI) containing antisense<br />

ODN that blocks the expression of c-raf gene was evaluated <strong>by</strong> intratumoral injection in nude<br />

mice with subcutaneously implanted solid tumor xenograft derived from human ovarian cancer. 18<br />

Significant retardation of solid tumor growth was observed in PEC formulation with antisense c-raf<br />

ODN compared to the formulations with a scrambled ODN and PBS (Figure 24.3). When the PEC<br />

micelles containing antisense c-raf ODN were intravenously administered to the mice bearing a<br />

Tumor volume (cm 3 )<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 2<br />

PEG miceless (antisense ODN)<br />

PEG miceless (mismatched ODN)<br />

Antisense ODN alone<br />

Control<br />

4 6<br />

8 10 12 14 16 18 20<br />

Time (Day)<br />

FIGURE 24.3 Effect of ODN–PEG/PEI PEC micelles containing antisense c-raf ODN on the growth of<br />

human ovarian cancer-derived tumor xenograft in nude mice. Each ODN formulation was administered<br />

intratumorally at 2.5 mg kg K1 injection K1 . (From Jeong, J. H., Kim, S. W., and Park, T. G., J. Control.<br />

Rel., 93, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


482<br />

Tumor volume (mm 3 )<br />

2000<br />

1500<br />

1000<br />

500<br />

Antisense MC<br />

Mismatched MC<br />

Control<br />

Antisense ODN Only<br />

0<br />

0 2 4 6<br />

Time (days)<br />

solid tumor derived from human lung cancer, the proliferation rate of the tumor was significantly<br />

reduced (Figure 24.4). 35 The PEC micellar formulation also exhibited much higher deposition of<br />

ODN in a solid tumor region than in a naked ODN after 12 h of intravenous injection through a tail<br />

vein (Figure 24.5). Taken together, the results suggested the hydrophilic PEG shell surrounding<br />

charge-neutralized polyelectrolyte core could successfully inhibit the nonspecific interaction of<br />

serum proteins in the blood stream. Also, through loosened vasculature near the rapidly growing<br />

tumors, the nanosized particles could be targeted to the solid tumor <strong>by</strong> the enhanced permeation and<br />

retention (EPR) effect. 43<br />

8<br />

10 12 14 16<br />

FIGURE 24.4 Effect of systemically administered ODN–PEG/PEI PEC micelles containing antisense c-raf<br />

ODN on the growth of human lung cancer-derived tumor xenograft in nude mice. Each ODN formulation was<br />

administered through tail vein at 2.5 mg kg K1 injection K1 . (From Jeong, J. H., Kim, S. H., Kim, S. W., and<br />

Park, T. G., Bioconjug. Chem., 16(4), 2005. With permission.)<br />

% ID/g tissue<br />

1.4<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

3 h<br />

12 h<br />

24 h<br />

ODN PEC Micelles<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 24.5 Accumulation of ODN–PEG/PEI PEC micelles in solid tumor (human lung cancer) after<br />

intravenous administration. (From Jeong, J. H., Kim, S. H., Kim, S. W., and Park, T. G., Bioconjug. Chem.,<br />

16(4), 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Targeted Antisense Oligonucleotide Micellar Delivery Systems 483<br />

24.3 LIGAND-MEDIATED TARGETING OF ANTISENSE OLIGONUCLEOTIDES<br />

Nonviral carriers do not have an inherent selectivity of cell surface binding and internalization. The<br />

limitation can be addressed <strong>by</strong> conjugating various ligands targeting specific cells or tissure to<br />

polymeric gene carriers. The PEC micelles demonstrated promising potential as a carrier for<br />

systemic cancer treatment that could target solid tumors. In conjunction with the passive targeting<br />

property, it would be advantageous if the PEC micelles have cell- or tissue-specific ligands on the<br />

surface. The specific ligand could play an additional role for enhanced internalization of PEC<br />

micelles into the cells <strong>by</strong> receptor mediated endocytosis.<br />

24.3.1 LIGAND-MEDIATED ODN TARGETING SYSTEMS BASED ON PEC MICELLES<br />

A graft copolymer composed of PEI and PEG where the terminal end of PEG was biotinylated for<br />

further conjugation was synthesized and used for ligand-mediated targeting of cancer cells. PEC<br />

micelles could be generated <strong>by</strong> interacting biotinylated PEI-g-PEG and ODN. Avidin was attached<br />

to the biotin located at the end of the PEG segment and biotinylated trasferrin (Tf) was subsequently<br />

attached to the avidin to obtain a PEI–PEG-Tf conjugate. 45 Enhanced cellular uptake of PEI–PEG-<br />

Tf/ODN PEC micelles was shown in cancer cells. When formulated with antisense ODN targeting<br />

multi-drug efflux transporter protein, P-glycoprotein (P-gp), the PEI–PEG-Tf/ODN PEC micelles<br />

efficiently inhibit the expression of the P-gp, leading to intracellular accumulation of a P-gp specific<br />

probe (rhodamine 123) in multi-drug resistant (MDR) cancer cells.<br />

Based on the same idea of PEG–ODN, a small interference lactosylated PEG-siRNA (Lac-<br />

PEG-siRNA) conjugate was recently synthesized. 46 An acid-labile b-thiopropionate linkage was<br />

introduced between siRNA and PEG to facilitate acid-triggered intracellular release of siRNA in<br />

the cytoplasm. The PEC micelles were prepared <strong>by</strong> interacting the PEG-siRNA conjugate with<br />

PLL, and they formed spherical shape nanoparticulates with an average diameter of. 117 nm. The<br />

lactose moiety at the end of the PEG segment could target the asialogycoprotein of hepatocytes to<br />

facilitate the receptor-mediated endocytosis of the PEC micelles. The PEC micelles demonstrated a<br />

significant RNA interference (RNAi) activity in human hepatoma cells (HuH-7), comparable to that<br />

of commercial cationic liposome-based carrier oligofectAMINE.<br />

24.3.2 IN VIVO APPLICATIONS<br />

PEC micelles demonstrated potential as a tumor targeting carrier for antisense ODN. Owing to their<br />

small size (generally less than 100 nm), the PEC micelles could be passively located in a solid<br />

tumor region through the loosened endothelial vasculature of the tumors <strong>by</strong> EPR effect. In addition<br />

to the passive targeting property, it would be more desirable if the carrier could also target a specific<br />

cell or tissue. Folate (FA) has been recognized as a potential targeting ligand for cancer cells.<br />

Expression of FA receptor is frequently up-regulated in a number of malignant tumors and highly<br />

restricted in normal cells and tissues. 47,48 Therefore, FA was coupled to several polymeric carriers<br />

for drug targeting to cancer cell. 38,49,50 FA has also been used for the modification of cationic<br />

polymers for tumor-targeted gene delivery. 51–53 Recently, ODN–PEG conjugate having FA at the<br />

end of PEG was synthesized for the systemic treatment of cancer based on PEC micelle delivery<br />

system. 54 The size of ODN–PEG–FA/PEI PEC micelles was ca. 90 nm with narrow distribution.<br />

Systemic administration of the micellar formulation to nude mice bearing human tumor xenograft<br />

derived from KB cells, FA-over-expressed the cells (KB cells), demonstrating the enhanced deposition<br />

of ODN in solid tumor region when compared to ODN–PEG/PEI PEC micelles. However, no<br />

significance in tumor deposition of ODN was observed between ODN–PEG–FA and ODN–PEG<br />

formulations in FA-receptor negative tumor xenograft model (A549 cells).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


484<br />

24.4 CONCLUSIONS<br />

Polymeric gene carriers have been developed as substitutes for viral vectors that could cause a<br />

series of unexpected side effects such as immunogenecity and oncogenecity. Micelle-based gene<br />

carriers exhibited a number of advantageous features for systemic delivery, including colloidal<br />

stability in solution, reduced interaction with components of biological fluid, improved in vivo<br />

pharmacokinetics, and the possibility of passive and active targeting of cancers. The intracellular<br />

pharmacokinetic behavior of the nucleic acid-based therapeutics could also be improved <strong>by</strong><br />

employing functional polymers with the existing formulation. Although there is still room for<br />

further improvement, the micellar gene delivery system could be considered a good example of<br />

a multi-functional drug delivery platform that has the potential to overcome a number of intrinsic<br />

barriers in the in vivo application of therapeutic genes.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Simons, M., Edelman, E. R., DeKeyser, J. L., Langer, R., and Rosenberg, R. D., Antisense c-myb<br />

oligonucleotides inhibit intimal arterial smooth muscle cell accumulation in vivo, Nature, 359, 67,<br />

1992.<br />

2. Stein, C. A. and Cheng, Y. C., Antisense oligonucleotides as therapeutic agents—is the bullet really<br />

magical?, Science, 261, 1004, 1993.<br />

3. Elbashir, S. M., Harborth, J., Lendeckel, W., Yalcin, A., Weber, K., and Tuschl, T., Duplexes of<br />

21-nucleotide RNAs mediate RNA interference in cultured mammalian cells, Nature, 411, 494, 2001.<br />

4. Elbashir, S. M., Lendeckel, W., and Tuschl, T., RNA interference is mediated <strong>by</strong> 21- and 22-nucleotide<br />

RNAs, Genes Dev., 15, 188, 2001.<br />

5. Wagner, R. W., Gene inhibition using antisense oligodeoxynucleotides, Nature, 372, 333, 1994.<br />

6. Jeong, J. H. and Park, T. G., Novel polymer-DNA hybrid polymeric micelles composed of hydrophobic<br />

poly(D,L-lactic-co-glycolic acid) and hydrophilic oligonucleotides, Bioconjug. Chem., 12, 917,<br />

2001.<br />

7. Braasch, D. A., Paroo, Z., Constantinescu, A., Ren, G., Oz, O. K., Mason, R. P., and Corey, D. R.,<br />

Biodistribution of phosphodiester and phosphorothioate siRNA, Bioorg. Med. Chem. Lett., 14, 1139,<br />

2004.<br />

8. Agrawal, S., Goodchild, J., Civeira, M. P., Thornton, A. H., Sarin, P. S., and Zamecnik, P. C.,<br />

Oligodeoxynucleoside phosphoramidates and phosphorothioates as inhibitors of human immunodeficiency<br />

virus, Proc. Natl. Acad. Sci. USA, 85, 7079, 1988.<br />

9. Nielsen, P. E., Egholm, M., Berg, R. H., and Buchardt, O., Peptide nucleic acids (PNAs): Potential<br />

antisense and anti-gene agents, Anticancer Drug Des., 8, 53, 1993.<br />

10. Egholm, M., Buchardt, O., Christensen, L., Behrens, C., Freier, S. M., Driver, D. A., Berg, R. H., Kim,<br />

S. K., Norden, B., and Nielsen, P. E., PNA hybridizes to complementary oligonucleotides obeying the<br />

Watson–Crick hydrogen-bonding rules, Nature, 365, 566, 1993.<br />

11. Pagnan, G., Stuart, D. D., Pastorino, F., Raffaghello, L., Montaldo, P. G., Allen, T. M., Calabretta, B.,<br />

and Ponzoni, M., Delivery of c-myb antisense oligodeoxynucleotides to human neuroblastoma cells<br />

via disialoganglioside GD(2)-targeted immunoliposomes: Antitumor effects, J. Natl. Cancer Inst., 92,<br />

253, 2000.<br />

12. Juliano, R. L. and Akhtar, S., Liposomes as a drug delivery system for antisense oligonucleotides,<br />

Antisense Res. Dev., 2, 165, 1992.<br />

13. Barron, L. G., Meyer, K. B., and Szoka, F. C. Jr., Effects of complement depletion on the pharmacokinetics<br />

and gene delivery mediated <strong>by</strong> cationic lipid-DNA complexes, Hum. Gene Ther., 9, 315,<br />

1998.<br />

14. Meyer, O., Kirpotin, D., Hong, K., Sternberg, B., Park, J. W., Woodle, M. C., and Papahadjopoulos,<br />

D., Cationic liposomes coated with polyethylene glycol as carriers for oligonucleotides, J. Biol.<br />

Chem., 273, 15621, 1998.<br />

15. Kim, J. S., Kim, B. I., Maruyama, A., Akaike, T., and Kim, S. W., A new non-viral DNA delivery<br />

vector: The terplex system, J. Control. Rel., 53, 175, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Targeted Antisense Oligonucleotide Micellar Delivery Systems 485<br />

16. Hashida, M., Takemura, S., Nishikawa, M., and Takakura, Y., Targeted delivery of plasmid DNA<br />

complexed with galactosylated poly(L-lysine), J. Control. Rel., 53, 301, 1998.<br />

17. Nishikawa, M., Takemura, S., Takakura, Y., and Hashida, M., Targeted delivery of plasmid DNA to<br />

hepatocytes in vivo: Optimization of the pharmacokinetics of plasmid DNA/galactosylated<br />

poly(L-lysine) complexes <strong>by</strong> controlling their physicochemical properties, J. Pharmacol. Exp.<br />

Ther., 287, 408, 1998.<br />

18. Kataoka, K., Togawa, H., Harada, A., Yasugi, K., Matsumoto, T., and Katayose, S., Spontaneous<br />

formation of polyion complex micelles with narrow distribution from antisense oligonucleotide and<br />

cationic block copolymer in physiological saline, Macromolecules, 29, 8556, 1996.<br />

19. Bronich, T. K., Cherry, T., Vinogradov, S. V., Eisenberg, A., Kabanov, V. A., and Kabanov, A. V.,<br />

Self-assembly in mixtures of poly(ethyleneoxide)-graft-poly(ethylenimine) and alkyl sulfates, Langmuir,<br />

14, 6101, 1998.<br />

20. Jeong, J. H., Kim, S. W., and Park, T. G., Novel intracellular delivery system of antisense oligonucleotide<br />

<strong>by</strong> self-assembled hybrid micelles composed of DNA/PEG conjugate and cationic fusogenic<br />

peptide, Bioconjug. Chem., 14, 473, 2003.<br />

21. Yoo, H. S. and Park, T. G., Biodegradable polymeric micelles composed of doxorubicin conjugated<br />

PLGA–PEG block copolymer, J. Control. Rel., 70, 63, 2001.<br />

22. Jeong, J. H., Song, S. H., Lim, D. W., Lee, H., and Park, T. G., DNA transfection using linear<br />

poly(ethylenimine) prepared <strong>by</strong> controlled acid hydrolysis of poly(2-ethyl-2-oxazoline), J. Control.<br />

Rel., 73, 391, 2001.<br />

23. Lee, H., Jeong, J. H., and Park, T. G., PEG grafted polylysine with fusogenic peptide for gene<br />

delivery: High transfection efficiency with low cytotoxicity, J. Control. Rel., 79, 283, 2002.<br />

24. Lee, H., Jeong, J. H., and Park, T. G., A new gene delivery formulation of polyethylenimine/DNA<br />

complexes coated with PEG conjugated fusogenic peptide, J. Control. Rel., 76, 183, 2001.<br />

25. Brus, C., Petersen, H., Aigner, A., Czubayko, F., and Kissel, T., Efficiency of polyethylenimines and<br />

polyethylenimine-graft-poly(ethylene glycol) block copolymers to protect oligonucleotides against<br />

enzymatic degradation, Eur. J. Pharm. Biopharm., 57, 427, 2004.<br />

26. Brus, C., Petersen, H., Aigner, A., Czubayko, F., and Kissel, T., Physicochemical and biological<br />

characterization of polyethylenimine-graft-poly(ethylene glycol) block copolymers as a delivery<br />

system for oligonucleotides and ribozymes, Bioconjug. Chem., 15, 677, 2004.<br />

27. Kunath, K., von Harpe, A., Petersen, H., Fischer, D., Voigt, K., Kissel, T., and Bickel, U., The<br />

structure of PEG-modified poly(ethylene imines) influences biodistribution and pharmacokinetics<br />

of their complexes with NF-kappaB decoy in mice, Pharm. Res., 19, 810, 2002.<br />

28. Jeong, J. H., Lee, M., Kim, W. J., Yockman, J. W., Park, T. G., Kim, Y. H., and Kim, S. W., Anti-GAD<br />

antibody targeted non-viral gene delivery to islet beta cells, J. Control. Rel., 107, 562, 2005.<br />

29. Katayose, S. and Kataoka, K., Water-soluble polyion complex associates of DNA and poly(ethylene<br />

glycol)-poly(1-lysine) block copolymer, Bioconjug. Chem., 8, 702, 1997.<br />

30. Katayose, S. and Kataoka, K., Remarkable increase in nuclease resistance of plasmid DNA through<br />

supramolecular assembly with poly(ethylene glycol)-poly(L-lysine) block copolymer, J. Pharm. Sci.,<br />

87, 160, 1998.<br />

31. Harada, A., Togawa, H., and Kataoka, K., Physicochemical properties and nuclease resistance of<br />

antisense–oligodeoxynucleotides entrapped in the core of polyion complex micelles composed of<br />

poly(ethylene glycol)-poly(L-lysine) block copolymers, Eur. J. Pharm. Sci., 13, 35, 2001.<br />

32. Kakizawa, Y., Harada, A., and Kataoka, K., Glutathione-sensitive stabilization of block copolymer<br />

micelles composed of antisense DNA and thiolated poly(ethylene glycol)-block-poly(L-lysine): A<br />

potential carrier for systemic delivery of antisense DNA, Biomacromolecules, 2, 491, 2001.<br />

33. Vinogradov, S. V., Bronich, T. K., and Kabanov, A. V., Self-assembly of polyamine-poly(ethylene<br />

glycol) copolymers with phosphorothioate oligonucleotides, Bioconjug. Chem., 9, 805, 1998.<br />

34. Kabanov, V. A. and Kabanov, A. V., Interpolyelectrolyte and block ionomer complexes for gene<br />

delivery: Physico-chemical aspects, Adv. Drug Deliv. Rev., 30, 49, 1998.<br />

35. Jeong, J. H., Kim, S. H., Kim, S. W., and Park, T. G., Polyelectrolyte complex micelles composed of<br />

c-raf anti-sense oligodeoxynucleotide-poly(ethylene glycol) conjugate and poly(ethylenimine): Effect<br />

of systemic administration on tumor growth, Bioconjug. Chem., 16(4), 1034–1037, 2005.<br />

36. Jeong, J. H., Kim, S. W., and Park, T. G., A new antisense oligonucleotide delivery system based on<br />

self-assembled ODN–PEG hybrid conjugate micelles, J. Control. Rel., 93, 183, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


486<br />

Nanotechnology for Cancer Therapy<br />

37. Oishi, M., Sasaki, S., Nagasaki, Y., and Kataoka, K., pH-responsive oligodeoxynucleotide (ODN)poly(ethylene<br />

glycol) conjugate through acid-labile beta-thiopropionate linkage: Preparation and<br />

polyion complex micelle formation, Biomacromolecules, 4, 1426, 2003.<br />

38. Yoo, H. S. and Park, T. G., Folate receptor targeted biodegradable polymeric doxorubicin micelles,<br />

J. Control. Rel., 96, 273, 2004.<br />

39. Yoo, H. S., Lee, E. A., and Park, T. G, Doxorubicin-conjugated biodegradable polymeric micelles<br />

having acid-cleavable linkages, J. Control. Rel., 82, 17, 2002.<br />

40. Harada-Shiba, M., Yamauchi, K., Harada, A., Takamisawa, I., Shimokado, K., and Kataoka, K.,<br />

Polyion complex micelles as vectors in gene therapy–pharmacokinetics and in vivo gene transfer,<br />

Gene Ther., 9, 407, 2002.<br />

41. Kakizawa, Y. and Kataoka, K., Block copolymer micelles for delivery of gene and related<br />

compounds, Adv. Drug Deliv. Rev., 54, 203, 2002.<br />

42. Ideta, R., Yanagi, Y., Tamaki, Y., Tasaka, F., Harada, A., and Kataoka, K., Effective accumulation of<br />

polyion complex micelle to experimental choroidal neovascularization in rats, FEBS Lett., 557, 21,<br />

2004.<br />

43. Maeda, H., Wu, J., Sawa, T., Matsumura, Y., and Hori, K., Tumor vascular permeability and the EPR<br />

effect in macromolecular therapeutics: A review, J. Control. Rel., 65, 271, 2000.<br />

44. Boussif, O., Lezoualc’h, F., Zanta, M. A., Mergny, M. D., Scherman, D., Demeneix, B., and Behr,<br />

J. P., A versatile vector for gene and oligonucleotide transfer into cells in culture and in vivo: Polyethylenimine,<br />

Proc. Natl. Acad. Sci. USA, 92, 7297, 1995.<br />

45. Vinogradov, S., Batrakova, E., Li, S., and Kabanov, A., Polyion complex micelles with proteinmodified<br />

corona for receptor-mediated delivery of oligonucleotides into cells, Bioconjug. Chem.,<br />

10, 851, 1999.<br />

46. Oishi, M., Nagasaki, Y., Itaka, K., Nishiyama, N., and Kataoka, K., Lactosylated poly(ethylene<br />

glycol)-siRNA conjugate through acid-labile beta-thiopropionate linkage to construct pH-sensitive<br />

polyion complex micelles achieving enhanced gene silencing in hepatoma cells, J. Am. Chem. Soc.,<br />

127, 1624, 2005.<br />

47. Lu, Y. and Low, P. S., Folate-mediated delivery of macromolecular anticancer therapeutic agents,<br />

Adv. Drug Deliv. Rev., 54, 675, 2002.<br />

48. Antony, A. C., The biological chemistry of folate receptors, Blood, 79, 2807, 1992.<br />

49. Kim, S. H., Jeong, J. H., Chun, K. W., and Park, T.G, Target-specific cellular uptake of PLGA<br />

nanoparticles coated with poly(1-lysine)-poly(ethylene glycol)-folate conjugate, Langmuir, 21,<br />

8852, 2005.<br />

50. Yoo, H. S. and Park, T. G., Folate-receptor-targeted delivery of doxorubicin nano-aggregates stabilized<br />

<strong>by</strong> doxorubicin-PEG-folate conjugate, J. Control. Rel., 100, 247, 2004.<br />

51. Cho, K. C., Kim, S. H., Jeong, J. H., and Park, T. G., Folate receptor-mediated gene delivery using<br />

folate-poly(ethylene glycol)-poly(L-lysine) conjugate, Macromol. Biosci., 5, 512, 2005.<br />

52. Benns, J. M., Maheshwari, A., Furgeson, D. Y., Mahato, R. I., and Kim, S. W., Folate-PEG-folategraft-polyethylenimine-based<br />

gene delivery, J. Drug Target., 9, 123, 2001.<br />

53. Kim, S. H., Jeong, J. H., Cho, K. C., Kim, S. W., and Park, T. G., Target-specific gene silencing <strong>by</strong><br />

siRNA plasmid DNA complexed with folate-modified poly(ethylenimine), J. Control. Rel., 104, 223,<br />

2005.<br />

54. Jeong, J. H., Kim, S. H., Kim, S. W., and Park, T. G., In vivo tumor targeting of ODN–PEG-folic<br />

acid/PEI polyelectrolyte complex micelles, J. Biomater. Sci. Polym. Ed., 16, 1409, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


25<br />

CONTENTS<br />

Dendrimers as Drug and Gene<br />

Delivery Systems<br />

Tae-il Kim and Jong-Sang Park<br />

25.1 Introduction ....................................................................................................................... 489<br />

25.2 Drug Delivery .................................................................................................................... 490<br />

25.2.1 Introduction ......................................................................................................... 490<br />

25.2.2 Non-Covalent Encapsulation of Drugs ............................................................... 490<br />

25.2.3 Covalent Conjugation of Drugs .......................................................................... 493<br />

25.2.4 Neutron Capture Therapy.................................................................................... 494<br />

25.2.4.1 Boron Neutron Capture Therapy ....................................................... 494<br />

25.2.4.2 Gadolinium Neutron Capture Therapy .............................................. 495<br />

25.2.5 Photodynamic Therapy ....................................................................................... 496<br />

25.3 Gene Delivery.................................................................................................................... 498<br />

25.3.1 Introduction ......................................................................................................... 498<br />

25.3.2 PAMAM Dendrimers .......................................................................................... 498<br />

25.3.3 Polypropylenimine Dendrimers .......................................................................... 501<br />

25.3.4 Other Dendrimers ................................................................................................ 502<br />

25.4 Conclusion ......................................................................................................................... 504<br />

Acknowledgments ........................................................................................................................ 504<br />

References..................................................................................................................................... 504<br />

25.1 INTRODUCTION<br />

Dendrimers are highly branched, mono-disperse macromolecules, characterized <strong>by</strong> layers from a<br />

core focal point called generations and multivalent end group functionalities. The word dendrimer<br />

originated from the Greek dendron, meaning “tree” and meros, meaning “part.”<br />

Dendrimers are prepared <strong>by</strong> iterative synthetic steps. First, Vögtle and coworkers reported the<br />

synthesis of dendritic structures as cascade molecules. 1 A lot of dendrimers have been developed,<br />

and they have become a subject of intense research, including within electrochemistry, photophysics,<br />

and supramolecular chemistry. 2,3 Over the past decade, characteristics such as a highly defined<br />

structure, modifiable surface functionality, and an internal cavity of dendrimers have aroused<br />

scientists’ interest in the biomedical field. In recent years, dendrimers have shown potential in<br />

fields, including drug delivery, gene delivery, magnetic resonance imaging (MRI), and anti-cancer<br />

therapeutics. 4–6<br />

The focus of this review is the history and the current trends of dendrimer application for drug<br />

and gene delivery systems.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

489


490<br />

25.2 DRUG DELIVERY<br />

25.2.1 INTRODUCTION<br />

The development of an efficient drug delivery system is very important to improve the pharmacological<br />

activity of drug molecules. Various types of drug delivery systems have been designed,<br />

including liposomes, polymeric micelles, etc. 7–9 Dendrimers have emerged as new alternatives and<br />

efficient tools for delivery of drug molecules. Figure 25.1 shows the molecular structures of various<br />

dendrimers used for delivery systems. Generally, bioactive drugs have hydrophobic moieties in<br />

their structure, resulting in low water-solubility that inhibits efficient delivery into cells. Dendrimers<br />

designed to be highly water-soluble and biocompatible have been shown to be able to improve<br />

drug properties such as solubility and plasma circulation time via various formulations and to<br />

deliver drugs efficiently. In comparison to linear polymeric carriers, the multivalent functionalities<br />

of dendrimers can be linked to drug molecules or ligands in a well-defined manner and can be used<br />

to increase the binding efficiency and affinity of therapeutic molecules to receptors via synergistic<br />

interaction. In addition, the low polydispersity of dendrimers can show a reproducible<br />

pharmacokinetic pattern.<br />

25.2.2 NON-COVALENT ENCAPSULATION OF DRUGS<br />

Drug delivery using unimolecular micelles has the advantage that the micellar structure is maintained<br />

and not disrupted at all concentrations, contrary to the behavior of conventional polymeric<br />

+ − − +<br />

− CO2 R4N<br />

+ NR4 − O2C<br />

+ −<br />

CO2 R4N NR4 O2C − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ NR4 − O2C<br />

+ −<br />

NR4 O2C<br />

+ −<br />

NR4 O2C<br />

CH3(OCH2CH2)16O<br />

CH3(OCH 2CH2)16O<br />

CH3(OCH 2CH2)16O<br />

CH 3(OCH2CH2)16O<br />

CH3(OCH2CH2)16O CH3(OCH2CH2)16O O(CH 2CH2O)16CH3<br />

O(CH 2CH2O)16CH3<br />

− +<br />

CO2 R4N<br />

CO2 − R4N +<br />

− +<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C<br />

CO2 R4N<br />

+ − − +<br />

NR4 O2C − + − CO2 R4N<br />

+NR4−<br />

CO2 R4N NR4 O2C − +<br />

O2C<br />

CO2 R4N<br />

(a) (b)<br />

O<br />

O<br />

O<br />

O<br />

O O<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O(CH2CH 2O)16CH3<br />

O(CH2CH2O)16CH3 O(CH2CH2O)16CH3 O(CH 2CH2O)16CH3<br />

(c) (d)<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

NH2 H2N<br />

NH HN<br />

O<br />

O<br />

N<br />

H2N O<br />

HN<br />

HN<br />

O<br />

N<br />

NH<br />

HN<br />

O O N O<br />

HN<br />

H2N N<br />

H2N<br />

NH<br />

O<br />

NH2<br />

HN<br />

N O<br />

NH2 O<br />

O NH NH<br />

H<br />

N N<br />

N<br />

O O NH<br />

NH<br />

O<br />

NH2 N<br />

HN<br />

H2N O<br />

O<br />

O<br />

HN<br />

N<br />

N<br />

N<br />

H<br />

HN O<br />

HN<br />

O<br />

H2N O N<br />

NH O<br />

H2N NH<br />

NH2 NH<br />

O<br />

NH2 N O O<br />

NH<br />

HN<br />

O N<br />

NH<br />

NH<br />

O<br />

N<br />

NH2 O O<br />

NH NH<br />

H2N NH2 HO<br />

HO<br />

HO<br />

HO<br />

HO<br />

HO<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

HO<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

O<br />

HO HO<br />

OH<br />

OH<br />

O O<br />

OH<br />

O O<br />

OH<br />

O<br />

O<br />

O<br />

O O O O<br />

O<br />

O<br />

O O<br />

FIGURE 25.1 Structures of various dendrimers. (a) Unimolecular micelle, (b) poly(amidoamine) dendrimer,<br />

(c) polyaryl ether dendrimer, (d) polyester dendrimer based on 2,2-bis(hydroxymethyl)propionic acid.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

OH<br />

OH<br />

O<br />

O<br />

O<br />

O<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH<br />

OH


Dendrimers as Drug and Gene Delivery Systems 491<br />

micelles because all segments are covalently connected. Newkome et al. reported that a dendritic<br />

unimolecular micelle was shown to solubilize various hydrophobic guest molecules in the internal<br />

hydrophobic cavities. 10 However, the problem of these systems for drug delivery is their high<br />

hydrophobicity. So, water-soluble and biocompatible poly(ethylene glycol) (PEG) has been introduced<br />

to the drug delivery systems.<br />

Initially, dendrimers were used for delivery systems mediated <strong>by</strong> non-covalent encapsulation of<br />

drugs as unimolecular micelles and dendritic boxes. Meijer and coworkers first reported the<br />

dendritic box based on the construction of a chiral shell of protected amino acids onto poly(propyleneimine)<br />

(PPI) dendrimers with 64 amine end groups (Figure 25.2). 11 They showed that the<br />

guest’s and cavity’s shapes determine the number of guests entrapped in the dendritic box; the<br />

architecture of the dendrimer is important, and shape-selective liberation of guests could be<br />

achieved <strong>by</strong> removing the shell in two steps. 12 Meijer and coworkers also developed a new<br />

host–guest system <strong>by</strong> synthesizing oligoethyleneoxy-modified PPI dendrimers, and they studied<br />

the interactions in aqueous media <strong>by</strong> UV/Vis titration. 13 PEGylated PPI dendrimers were also<br />

prepared as drug delivery systems <strong>by</strong> Pales and coworkers. 14 PEGylated PPI dendrimers were<br />

O<br />

O O<br />

o<br />

N<br />

N<br />

H<br />

N<br />

N<br />

N<br />

H<br />

N<br />

N<br />

H<br />

N N<br />

H<br />

N<br />

N<br />

N<br />

H<br />

H<br />

N<br />

N<br />

N<br />

H<br />

N<br />

H<br />

N N<br />

N<br />

N<br />

H<br />

N<br />

N<br />

N<br />

H<br />

N<br />

N<br />

R O<br />

O<br />

O<br />

O O<br />

O<br />

N<br />

O<br />

O<br />

N<br />

O O<br />

N<br />

N<br />

O<br />

N<br />

O<br />

N<br />

N O<br />

N N<br />

N O O<br />

O<br />

N N<br />

N<br />

O N<br />

N<br />

N O<br />

N<br />

O<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

N<br />

O<br />

O O O<br />

O O<br />

NH HN<br />

O O<br />

NH HN<br />

O NH<br />

O<br />

O<br />

HN NH<br />

NH HN<br />

O<br />

O<br />

O<br />

NH<br />

HN<br />

O<br />

HN<br />

O O<br />

O HN<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HN<br />

HN<br />

NH<br />

HN<br />

NH HN<br />

HN NH<br />

NH<br />

NH HN<br />

O O<br />

HN<br />

HN NH<br />

NH HN<br />

HN<br />

NH<br />

NH<br />

H<br />

NH<br />

H<br />

H<br />

H<br />

H<br />

H<br />

H<br />

H<br />

H<br />

NH<br />

H<br />

NH<br />

NH<br />

HN<br />

NH HN<br />

HN NH<br />

HN<br />

NH HN<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

OO<br />

O O<br />

O<br />

O O<br />

O<br />

R O R<br />

O<br />

R<br />

R<br />

R<br />

R<br />

RO<br />

R<br />

R<br />

RO<br />

R<br />

R<br />

R<br />

R<br />

R R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

R<br />

O<br />

O<br />

R<br />

O<br />

O O<br />

R<br />

O<br />

O<br />

R<br />

O O<br />

O<br />

R O<br />

O<br />

R OO R O<br />

O O<br />

O O<br />

Guest molecule<br />

FIGURE 25.2 Scheme of modified PPI dendrimers for the non-covalent encapsulation of guest molecules.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

R


492<br />

Nanotechnology for Cancer Therapy<br />

shown to enhance water-solubility of encapsulated pyrene and betamethasones, securing their<br />

application as promising controlled release drug carriers. Encapsulated molecules were found to<br />

be solubilized in both regions of the interior of dendrimers and the PEG coat. The same group<br />

reported that the introduction of quaternary ammonium groups to PPI dendrimers not only<br />

enhanced the water solubility, but they also affected the protonation titration profile, leading to<br />

the release of the entrapped molecule within a narrower pH region. 15 Pyrene is released when the<br />

internal tertiary amines get protonated between pH 4 and 2, suggesting these dendrimers are<br />

potential candidates for pH-sensitive controlled-release drug delivery systems.<br />

Poly(amidoamine) (PAMAM) dendrimer that was first developed <strong>by</strong> Tomalia et al. 16 also has<br />

been extensively studied for non-covalent encapsulation of drug molecules. PAMAM dendrimers<br />

are water-soluble and biocompatible, and they possess internal cavities hosting guest molecules.<br />

Twyman et al. modified the ester end groups of half-generation PAMAM dendrimers with Tris. 17<br />

This highly water-soluble dendrimer could solubilize hydrophobic guest molecules at pH 7.0, but it<br />

lost its binding capability to guest molecules at pH 2.0, presumably because of the protonation of<br />

the internal tertiary amines.<br />

To improve biocompatibility and water-solubility, PEG was grafted on the surface of PAMAM<br />

dendrimers. PEG-attached PAMAM dendrimers could encapsulate anti-cancer drugs, adriamycin,<br />

and methotrexate, and their ability to encapsulate these drugs increased with the increasing<br />

dendrimer generation and chain length of PEG grafts. 18 The dendrimers slowly released the<br />

drugs in low ionic strength solution; however, drugs were readily released in isotonic solutions.<br />

Another group explored the use of PEGylated PAMAM dendrimers for delivery of the anti-cancer<br />

drug 5-fluorouracil (5-FU) in vitro and in albino rats. 19 PEGylation of dendrimers was found to<br />

increase their drug-loading capacity and reduce the drug release rate and hemolytic toxicity,<br />

suggesting the use of PEGylated PAMAM dendrimers as prolonged delivery systems.<br />

Haba et al. prepared PEG-modified PAMAM dendrimers having methacryloyl groups on the<br />

chain ends of the dendrimer for drug delivery. 20 After linking of the methacryloyl groups, they<br />

found that the dendrimer decreased its affinity against rose Bengal because of its hiding the<br />

dendrimer interior of the peripheral network and that the dendrimer having the peripheral<br />

network retained Rose Bengal dye molecules tightly in their interior. This result indicated that<br />

the linking of polymerizable groups attached to the periphery of the dendrimer could be an efficient<br />

approach for retaining guest molecules in their interior, and that they could be used as photosensitizers<br />

for photodynamic therapy (PDT) and as boron10-containing compounds for neutron<br />

capture therapy because these pharmaceutical molecules could exhibit their therapeutic effects,<br />

even if they are not released from their carriers.<br />

Frechet and coworkers synthesized novel dendritic unimolecular micelles with a hydrophobic<br />

polyether dendrimer core surrounded <strong>by</strong> hydrophilic PEGs. 21 For the G-3 micelle, the loading of<br />

indomethacin was found to be of 11%, a value corresponding to approximately nine to ten drug<br />

molecules per micelle, and preliminary in vitro release tests showed that sustained release characteristics<br />

were achieved. Recently, linear dendritic block copolymers comprising PEG and either a<br />

polylysine or polyester dendrons were designed, and highly acid-sensitive cyclic acetals were<br />

attached to the dendrimer periphery for pH-responsive micelle systems. 22 At pH 7.4, the fluorescence<br />

of micelle-encapsulated Nile Red was constant, indicating it was retained in the micelle,<br />

whereas at pH 5, the fluorescence decreased, consistent with its release into aqueous solution. The<br />

rate of release was strongly correlated with the rate of acetal hydrolysis. These results suggest that<br />

the loading and controlled-release of drugs are dependent on the chemical structure of<br />

the dendrimer.<br />

A novel triblock copolymer composed of citric acid dendritic blocks and PEG core was<br />

synthesized and applied as a drug delivery system. 23 The dendrimers were found to be able to<br />

trap various hydrophobic drugs in their suitable sites and to release them in a controlled manner in<br />

in vitro experiments.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 493<br />

25.2.3 COVALENT CONJUGATION OF DRUGS<br />

Another approach for dendritic drug carriers is based on covalent conjugation of drug molecules to<br />

the dendrimer periphery using the multivalency of dendrimers. Although many drug molecules<br />

were conjugated to the dendrimers through non-degradable linkages, introducing degradable bonds<br />

between the drug and the dendrimer can control the release of drug molecules. Figure 25.3 shows<br />

the scheme of drug-conjugated dendrimers.<br />

Dendrimers were first used for immunoconjugates. Chemical modification of the antibody with<br />

drug molecules often diminished its biological activity, but approaches to link the antibody to<br />

another macromolecule including the dendrimer were reported to increase drug loading while<br />

retaining the activity. Roberts et al. conjugated antibody to porphyrin to chelate radiolabeled<br />

copper ions using PAMAM dendrimer as a linker. 24 They found that 100% of the porphyrin<br />

bound to the heavy chain of the antibody, and dendrimer conjugates contributed 90% of the<br />

immunoactivity of unmodified antibody and showed a high radiolabeling level in contrast to the<br />

conjugates obtained through the random-coupling method. PAMAM dendrimers were also<br />

modified chemically <strong>by</strong> reaction with 1,4,7,l0-tetraazacyclododecanetetraacetic acid (DOTA)<br />

and diethylenetriaminepentaacetic acid (DTPA) type bifunctional metal chelators and used as a<br />

linker to monoclonal antibody (moAb) 2E4. 25 The dendrimer–antibody conjugates were found to<br />

retain biological activity and to be easily radiolabeled, suggesting use of these dendrimer conjugates<br />

for moAb-based radiotherapy or imaging. Recently, Patri et al. conjugated anti-PSMA<br />

(prostate-specific membrane antigen) to PAMAM dendrimer labeled with FITC and capped with<br />

acetic anhydride to minimize the non-specific interaction with the cell surfaces for targeting<br />

prostate cancer. 26 This dendrimer conjugate showed great potential to target PSMA-positive<br />

LNCaP cell line with minimal loss of immunoreactivity. Meanwhile, PAMAM dendrimers were<br />

utilized to investigate the potential for cancer chemotherapy <strong>by</strong> conjugating with anti-cancer drugs.<br />

Cisplatin, a potent anti-cancer drug, was conjugated to PAMAM dendrimer, giving a dendrimer–<br />

platinate that was highly water-soluble and released platinum slowly in vitro. 27,28 This dendrimer<br />

conjugate showed more potent anti-tumor activity than cisplatin and selective accumulation in solid<br />

tumor tissue and was also less toxic than cisplatin. Zhuo et al. synthesized a water-soluble acetylated<br />

PAMAM dendrimer series with a core of 1,4,7,10-tetraazacyclododecane and linked it to<br />

1-bromoacetyl-5-fluorouracil to form dendrimer–5-FU conjugates. 29 Hydrolysis of the conjugates<br />

in a phosphate buffer solution could release free 5-FU in a generation-dependent manner, showing<br />

the dendrimer conjugate seems to be a promising carrier for the controlled release of anti-tumor<br />

drugs. Wang et al. used PAMAM dendrimers for doxorubicin (DOX) delivery. 30 They conjugated<br />

semitelechelic poly[N-(2-hydroxypropyl)-methacrylamide] macromolecules (ST-PHPMA,) to the<br />

periphery of PAMAM dendrimer and introduced DOX to the dendrimer conjugate, star-like HPMA<br />

copolymer. It showed lower cytotoxicity than other brush-like polymer-DOX in A2780 human<br />

ovarian carcinoma cell line and a slow rate of in vitro DOX release, demonstrating that the<br />

differences of DOX internalization because of the structure of the polymers would result in different<br />

Drug<br />

Drug<br />

Targeting molecule<br />

FIGURE 25.3 Schematic view of drug–dendrimer conjugate. A dotted line, degradable linkage; a black line,<br />

non-degradable linkage.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


494<br />

intracellular DOX concentrations with cytotoxicity changes. 5-Aminosalicylic acid (5-ASA) was<br />

also conjugated to PAMAM dendrimers using two different spacers containing azo-bond, p-aminobenzoic<br />

acid (PABA), and p-aminohippuric acid (PAH) for colonic delivery. 31 The dendrimer<br />

conjugates incubated in rat fecal contents showed colon-specific and prolonged release of<br />

5-ASA through reductive cleavage of azo-bond <strong>by</strong>colonic bacteria, suggesting the potential for<br />

colon-specific drug delivery carriers. In addition, Quintana et al. attached folic acid to acetamide-,<br />

hydroxyl-, or carboxyl-capped PAMAM dendrimers conjugated with methotrexate (MTX) to target<br />

tumor cells through the folate receptor. 32 These dendrimer conjugates showed enhanced cellular<br />

uptake, and this targeted delivery improved the cytotoxic response of the cells to MTX 100-fold<br />

over the free drug. Recently, Tansey et al. also conjugated folic acid and indocyanine green, a<br />

model diagnostic agent to biodegradable poly(L-glutamic acid) (PG) chain-terminated PAMAM<br />

dendrimer for tumor targeting. 33 The PG–dendrimer conjugates showed a more prolonged<br />

degradation pattern than linear PG chain in the presence of lysosomal enzyme. After conjugation<br />

of PEG, the dendrimer conjugates were found to show reduced non-specific interaction and to bind<br />

selectively to tumor cells expressing folate receptors.<br />

Folic acids have already been conjugated to acid-labile hydrazide-terminated polyaryl ether<br />

dendrimers to target tumor cells <strong>by</strong> Frechet and coworkers. 34 The conjugates are soluble in aqueous<br />

medium above pH 7.4. The anti-tumor drug methotrexate was also attached to the dendrimers. The<br />

same group also prepared polyaryl ether dendrimer–PEG conjugates as model drug carriers. 35<br />

Cholesterol and two amino acid derivatives were attached to the dendrimer via carbonate, ester,<br />

and carbamate linkages.<br />

Greenwald and coworkers synthesized novel PEG–poly(aspartic) acid dendron conjugates and<br />

attached an anti-tumor agent, cytosine arabinoside (ara-C), to the dendron conjugates. 36 These<br />

conjugates showed greater tumor inhibition than free drugs and increased aqueous solubility,<br />

resulting in high loading and controlled release of drugs. Another study of conjugation of ara-C<br />

to PEG-dendrons composed of aminoadipic acid was accomplished <strong>by</strong> Schiavon et al. 37 The<br />

conjugates also presented prolonged blood residence time and higher loading of drugs.<br />

Polyester dendrimers based on the monomer unit 2,2-bis(hydroxymethyl)propanoic acid were<br />

also developed as drug carriers both in vitro and in vivo. 38 These dendrimers were found to be both<br />

water soluble and non-toxic. A 3-arm PEG–polyester dendrimer hybrid that conjugated with DOX<br />

via a hydrazone linkage susceptible to acidic hydrolysis showed pH-dependent drug release in vitro<br />

and a degree of anti-cancer activity in cancer cell lines. In addition, the little accumulation of the<br />

dendrimer conjugate found in any organ examined, including the liver, heart, and lungs, suggested<br />

that it is highly water soluble and biocompatible.<br />

Recently, a novel approach to developing covalently conjugated dendritic drug carriers was<br />

presented <strong>by</strong> two different groups. 39,40 These cascade-release or self-immolative conjugates were<br />

able to simultaneously release all drug molecules connected to the dendritic termini <strong>by</strong> a single<br />

trigger, resulting in whole dendrimer collapse. Moreover, Shabat and coworkers demonstrated selfimmolative<br />

dendrimers as a platform for multi-prodrugs, activated <strong>by</strong> a single enzymatic cleavage.<br />

41 The bioactivation of the heterodendritic prodrugs of DOX and campothecin (CPT) was<br />

evaluated using the Molt-3 leukemia cell line, and the conjugates showed a mild to significant<br />

increase in toxicity in comparison with the classical monomeric prodrugs.<br />

25.2.4 NEUTRON CAPTURE THERAPY<br />

25.2.4.1 Boron Neutron Capture Therapy<br />

Nanotechnology for Cancer Therapy<br />

Boron neutron capture therapy (BNCT) is a cancer-therapeutic application based on a nuclear<br />

capture reaction of a lethal 10 B(n,a) 7 Li 3C reaction. 42 If 10 B can be delivered at a concentration<br />

of at least 10 9 atoms per cell to tumor tissue, subsequent irradiation with thermal or epithermal<br />

neutrons produces highly energetic a particles and Li 3C ions that damage the tumor cells in nuclear<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 495<br />

fission reaction. To deliver high levels of boron in tumor tissue, it is proposed to use dendrimers as<br />

boron carriers for their well-defined molecular structure and multivalency.<br />

Initially, Barth et al. utilized PAMAM dendrimers in order to boronate moAb IB16-6 that<br />

targeted murine B16 melanoma using isocyanato polyhedral borane (Na(CH3)3NB10H8NCO). 43<br />

The boronated dendrimers were found to localize in the liver and spleen in in vivo distribution<br />

study, leaving a question of if the properties of boronated dendrimers can be modified to reduce<br />

their hepatic and splenic localization. Calssonn and coworkers attached epithermal growth factor<br />

(EGF) to boronated PAMAM dendrimers in order to target EGF receptor (EGFR) that is over<br />

expressed in brain tumors. 44 The boronated dendrimers-EGF were bound to malignant glioma cell<br />

membrane and endocytosed, resulting in the accumulation of boron in lysosome. Although these<br />

boronated dendrimers were localized in the liver and accumulated a little in the tumor after<br />

intravenous injection into rats having intracerebral C6 glioma transfected with EGFR gene,<br />

direct intratumoral injection could selectively deliver them to EGFR-positive gliomas. 45 Recently,<br />

use of another moAb that is directed against EGFR and EGFRvII, cetuximab (IMC-C225), was<br />

investigated for boronated PAMAM dendrimer-mediated BNCT. 46 There were w1100 boron<br />

atoms per molecule of cetuximab with only a slight reduction of Ka. Localization study of these<br />

conjugates in F98 rats bearing intracerebral implants of either EGFR-transfected or wild-type<br />

gliomas <strong>by</strong> intratumoral injection showed specific molecular targeting of EGFR and accumulation<br />

of boron.<br />

To reduce the hepatic uptake of boronated dendrimers, PEG was introduced to the dendrimers.<br />

47 Moreover, folic acid was also conjugated to the dendrimers to target the folate receptors in<br />

cancer cells. Among all the prepared combinations, boronated dendrimers with 1–1.5 PEG2000<br />

units exhibited the lowest hepatic uptake in C57BL/6 mice. Interestingly, dendrimers containing<br />

w13 decaborate clusters, w1 PEG2000 unit, and w1 PEG800 unit with folic acid attached to the<br />

distal end showed receptor-dependent uptake in contrast to dendrimers containing w15 decaborate<br />

clusters and w1 PEG2000 unit with folic acid attached to the distal end in in vitro studies using<br />

folate receptor (C) KB cells. Biodistribution studies with former dendrimer in C57BL/6 mice<br />

bearing folate receptor (C) murine 24JK-FBP sarcomas resulted in selective tumor uptake<br />

(6.0% ID/g tumor), and they also showed an increase in uptake of the dendrimers <strong>by</strong> the liver<br />

and kidney, warranting further effort in the optimization of biodistribution profiles of these<br />

boronated dendrimers.<br />

Polylysine dendrons also were examined for BNCT. Borane moieties were attached to the<br />

periphery of the dendrons and peptide spacers linked the dendrons to targeting molecules such<br />

as antibodies, a fluorescent moiety, and a PEG chain (Figure 25.4). 48 This offered a promising<br />

perspective for application in BNCT.<br />

25.2.4.2 Gadolinium Neutron Capture Therapy<br />

Although the vast majority of NCT research has involved various aspects of boron chemistry for<br />

BNCT, 157 Gd has also been suggested from theoretical and experimental studies with limited<br />

success. 49–51 To some extent, Gd complexes have been clinically employed as an MRI contrasting<br />

agent. 52 The potential to trace the in vivo pharmacokinetics of a Gd-NCT agent directly <strong>by</strong> MRI<br />

would be useful for planning the neutron irradiation and to predict the therapeutic effect.<br />

Kobayashi et al. synthesized avidin-G6-(1B4M-Gd)254(Av-G6Gd) from PAMAM dendrimer<br />

G6, biotin, avidin, and 2-(p-isothiocyanatobenzyl)-6-methyl-diethylenetriaminepentaacetic acid<br />

(1B4M). 53 A sufficient amount (162 ppm) of Av-G6Gd was accumulated and internalized into<br />

the SHIN3 cells both in vitro and in vivo to kill the cell using 157/155 Gd with external irradiation<br />

with an appropriate neutron beam while monitoring with MRI, suggesting that Av-G6Gd may be a<br />

promising agent for Gd neutron capture therapy of peritoneal carcinomatosis, and it may have the<br />

potential to permit monitoring of its pharmacokinetic progress with MRI.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


496<br />

H 2 N<br />

H 2 N<br />

N<br />

H<br />

O<br />

H N 2<br />

N<br />

H<br />

NH<br />

25.2.5 PHOTODYNAMIC THERAPY<br />

O<br />

O<br />

NH<br />

O<br />

NH 2<br />

HN<br />

O<br />

O<br />

O<br />

NH<br />

O O<br />

O<br />

O<br />

H<br />

H<br />

H<br />

N<br />

N<br />

N<br />

N<br />

N<br />

H<br />

H<br />

O O O<br />

HN<br />

NH 2<br />

HN<br />

O<br />

HN<br />

O<br />

NH 2<br />

O<br />

NH<br />

PDT is a technique for inducing tissue damage with light, following administration of a lightactivated<br />

photosensitizing drug that can be selectively concentrated in malignant or diseased<br />

tissue. 54 Activation of the photosensitizer results in the generation of reactive oxygen species<br />

(ROS), primarily singlet oxygen that oxidizes intracellular substrates such as lipids and amino<br />

acid residues, leading to cell death. The most widely studied application of PDT to date is in cancer<br />

therapy. However, the use of conventional photosensitizer systems bring skin phototoxicity,<br />

damage to normal tissue because of poor tumor selectivity, poor water solubility, and difficulties<br />

in treating solid tumors. In order to improve the disadvantages, dendrimers would be manufactured<br />

as promising photosensitizer carriers through modification of peripheral functionality.<br />

Battah et al. synthesized dendrimers with 5-aminolevulinic acid (ALA) moieties conjugated to<br />

the periphery <strong>by</strong> ester bonds that are cleavable <strong>by</strong> esterase inside the cells for PDT. 55 ALA is a<br />

natural precursor of an effective photosensitizer protophorphyrin IX (PpIX), and its administration<br />

can increased the cellular concentration of PpIX. 56 These ALA dendrimers showed increased<br />

NH 2<br />

N<br />

H<br />

NH 2<br />

NH 2<br />

NH 2<br />

Nanotechnology for Cancer Therapy<br />

O<br />

N<br />

H<br />

SH<br />

O<br />

HN SO 2<br />

O PEG<br />

FIGURE 25.4 Polylysine dendrimer conjugate used in boron neutron capture therapy (BNCT).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

N (CH 3 ) 2


Dendrimers as Drug and Gene Delivery Systems 497<br />

production of PpIX and higher cytotoxicity after irradiation relative to free ALA in tumorigenic<br />

keratinocyte PAM 212 cell line.<br />

Another approach of designing dendritic photosensitizers for PDT involved aryl ether<br />

dendrimer porphyrins (DP) with either 32 quaternary ammonium groups or carboxylic acid<br />

moieties on their periphery (Figure 25.5). 57 Hydrophilic DPs were found to be localized in<br />

membrane-limited organelles but hydrophobic PIX diffused through the cytoplasm, except the<br />

nucleus. DPs showed remarkably higher 1 O2-induced cytotoxicity against Lewis lung carcinoma<br />

(<strong>LLC</strong>) cells and far lower dark toxicity than PIX, demonstrating their highly selective photosensitizing<br />

effect in combination with a reduced systemic toxicity.<br />

Zhang et al. entrapped DP with 32 amine groups on the periphery, [NH2CH2CH2NHCO]32DPZn<br />

with polyion micells composed of PEG and poly(aspartic) acid for effective PDT. 58 Compared to<br />

[NH2CH2CH2NHCO]32DPZn, a relatively low cellular uptake of [NH2CH2CH2NHCO]32DPZn<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

X<br />

O<br />

O<br />

X<br />

O O<br />

X<br />

O<br />

O<br />

O O<br />

O<br />

X X<br />

O O<br />

X<br />

X = CONH(CH 2) 2N + Me 3CI − or COO − H +<br />

FIGURE 25.5 Porphyrin-based dendrimer for photodynamic therapy.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

N<br />

N<br />

Zn<br />

N<br />

N<br />

O<br />

X<br />

O<br />

O<br />

O<br />

X<br />

O O<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

X<br />

O O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X


498<br />

incorporated in the PIC micelle was observed, yet the latter exhibited enhanced photodynamic<br />

efficacy on the <strong>LLC</strong> cell line. Also, this system was found to reduce the dark toxicity of the cationic<br />

dendrimer porphyrin, probably as a result of the biocompatible PEG shell of the micelles.<br />

Another approach for deeper tissue penetration was tried using two-photon absorption (TPA)<br />

with near-infrared lasers. Porphyrin sensitizers have low TPA cross-sections, limiting their usefulness<br />

in TPA applications. Recently, Dichtel et al. attached donor chromophores capable of efficient<br />

TPA to a central porphyrin acceptor for an enhanced TPA cross-section. 59 This dendrimer was<br />

found to enhance the effective TPA cross-section of a porphyrin acceptor, allowing more efficient<br />

generation of singlet oxygen using 780 nm light. This result showed a potential use of PDT to<br />

tumors below the skin surface.<br />

25.3 GENE DELIVERY<br />

25.3.1 INTRODUCTION<br />

Cationic dendrimers have been reported to deliver genetic materials into cells <strong>by</strong> forming<br />

complexes with negatively charged genetic materials through electrostatic interaction. The<br />

complexes bound to cell membranes were known to be internalized into cells <strong>by</strong> endocytosis<br />

(Figure 25.6). It has been postulated that the branched cationic polymers such as polyethyleneimine<br />

or PAMAM dendrimer, having a high buffer capacity lead to polymeric swelling in acidic endosomes,<br />

disrupting the membrane barrier and resulting in release of the complexes. 60 However, the<br />

uptake and the transfection mechanism of cationic polymers are not yet fully understood. Recently,<br />

two different mechanisms, caveolae and clathrin-mediated endocytosis, were also reported to<br />

involve the uptake mechanism of the polyplexes. 61<br />

Several dendrimers have been synthesized and used for gene delivery carriers from a decade<br />

ago. The history is not long, but the intense research on developing dendritic gene delivery carriers<br />

has been carried out because of many characteristic advantages of dendrimers such as monodisperse<br />

size of dendrimers and controllable multivalent end-group functionalities.<br />

25.3.2 PAMAM DENDRIMERS<br />

It is only for a decade that the potential of dendrimers as gene delivery carriers has been investigated.<br />

The first applied dendrimer for gene delivery was PAMAM dendrimer that was reported <strong>by</strong><br />

Haensler and Szoka. 62 They showed that the dendrimers mediate high efficiency transfection of a<br />

variety of suspension- and adherent-cultured mammalian cells, and dendrimer-mediated<br />

DNA<br />

Dendrimer<br />

DNA-dendrimer<br />

complex<br />

Endocytosis<br />

FIGURE 25.6 Scheme of dendrimer-mediated transfection.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nucleus<br />

Endosome<br />

Nanotechnology for Cancer Therapy<br />

DNA<br />

transcription<br />

Protein<br />

translation


Dendrimers as Drug and Gene Delivery Systems 499<br />

transfection is a function both of the dendrimer/DNA ratio and the diameter of the dendrimer. When<br />

GALA, a water soluble, membrane-destabilizing peptide, is covalently attached to the dendrimer<br />

via a disulfide linkage, transfection efficiency is increased <strong>by</strong> 2–3 orders of magnitude. Also, they<br />

concluded that the high transfection efficiency of the dendrimers may be caused <strong>by</strong> the low pKa of<br />

the amines in the polymer that permit the dendrimer to buffer the pH change in the endosomal<br />

compartment. Kukowska-Latallo et al. also reported a high degree of transfection of PAMAM<br />

dendrimers with minimal cytotoxicity in various cell lines. 63 The addition of DEAE-dextran<br />

increased the transfection efficiency, suggesting that dextran alters the nature of dendrimer/DNA<br />

complexes. The most efficient transfection was achieved with a lysosomotropic agent such as<br />

chloroquine, indicating that endosomal localization is a rate-determining step in transfection.<br />

Mazda and coworkers investigated the transfection of the Epstein–Barr virus (EBV)-based<br />

plasmid vector/PAMAM dendrimer complexes to improve the efficiency in hepatocellular carcinoma<br />

cells (HCC) and cholangiocarcinoma cells. 64,65 They found highly efficient suicide gene<br />

transfer into various HCC cells <strong>by</strong> EBV-based plasmid vectors/PAMAM dendrimer complexes<br />

in vitro, suggesting a possible application of this non-viral vector system to gene therapy of HCC,<br />

and they reported that EBV-based plasmid vectors/PAMAM dendrimer complexes equipped with<br />

the carcinoembryonic antigen (CEA) promoter showed high transfection in cholangiocarcinoma<br />

cells, providing an efficient non-viral method for the targeted gene therapy of CEAproducing<br />

malignancies.<br />

Thermally degraded PAMAM dendrimers with heterodisperse structure (fractured PAMAM)<br />

showed a dramatically enhanced transfection efficiency in in vitro experiments. 66 Transfection<br />

activity was found to be related both to the initial size of the dendrimer and to its degree of<br />

degradation. The increased transfection after the heating process was thought to be principally<br />

due to the increase in flexibility that enabled the fractured dendrimer to be compact when<br />

complexed with DNA and swell when released from DNA. The dendrimers are presently sold<br />

commercially <strong>by</strong> Qiagen as SuperFecte.<br />

Bielinska et al. investigated the ability of PAMAM dendrimers to function as an effective<br />

delivery system for antisense oligonucleotides (ODNs) and antisense expression plasmids for the<br />

targeted modulation of gene expression. 67 Transfections of ODNs or antisense cDNA plasmids<br />

usingPAMAMdendrimersresultedinaspecificand dose-dependent inhibition of luciferase<br />

expression. Binding of the phosphodiester oligonucleotides to dendrimers also extended their<br />

intracellular survival. Delong et al. also evaluated PAMAM dendrimer (generation 3, Mr 6909)<br />

as a potential delivery vehicle for oligonucleotides. 68 Use of dendrimers resulted in a 50-fold<br />

enhancement in cell uptake of oligonucleotide as determined <strong>by</strong> flow cytometry, enhanced cytosolic,<br />

and nuclear availability as shown <strong>by</strong> confocal microscopy. Interestingly, fluorescent dye,<br />

Oregon green 488-conjugated dendrimer, was proved to be a much better delivery agent for antisense<br />

compounds than unmodified dendrimer. 69 This result suggested that coupling of relatively<br />

hydrophobic small molecules to PAMAM dendrimers may provide a useful means of enhancing<br />

their capabilities as delivery agents for nucleic acids.<br />

The efficiency of cell transfection <strong>by</strong> PAMAM dendrimers was shown to be greatly increased<br />

when anionic oligomers (oligonucleotide or dextran sulfate) were mixed with plasmids before<br />

addition of the dendrimer. 70 This enhancement of gene transfer <strong>by</strong> anionic oligomers depended<br />

on their size, structure, and charge. This result may be due to modification of the interaction<br />

between plasmids and dendrimers induced <strong>by</strong> the oligonucleotide that effect leads to a less<br />

condensed structure of the polyplexes, there<strong>by</strong> facilitating their penetration into cells.<br />

The transfection efficiency of PAMAM dendrimers was also increased <strong>by</strong> addition of substituted<br />

b-cyclodextrins (b-CD). 71 In vitro CAT expression was increased approximately 200-fold<br />

when dendrimer/DNA/b-CD formulations were applied on the surface of collagen membranes,<br />

indicating that b-CDs affected the physicochemical properties of dendrimer/DNA complexes.<br />

Besides, PAMAM dendrimers conjugated with a-cyclodextrins (a-CDs) were investigated for<br />

their potential as a gene delivery system. 72 The gene transfer activity of a-CD dendrimer conjugate<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


500<br />

Nanotechnology for Cancer Therapy<br />

(G3) was higher than that of dendrimers (G2, G3, and G4) and of the a-CD conjugate (G2), and the<br />

cytotoxicity of the a-CD conjugate (G3) was lower than that of the a-CD conjugate (G4). This<br />

greatest gene transfer activity could be attributed to the synergistic interplay of the proton sponge<br />

effect of the dendrimer (G3) and the tentative membrane-disrupting effect of a-CD on endosomal<br />

membranes. Recently, the transfection efficiency of a-CD dendrimer conjugates was improved <strong>by</strong><br />

surface modification of dendrimers with mannose. 73 The mannosylated conjugates were found to<br />

have a much higher gene transfer activity than unmodified a-CD dendrimer in various cell lines,<br />

which was independent of the expression of cell surface mannose receptors. Also, the conjugates<br />

provided higher gene transfer activity than dendrimer and a-CD dendrimer conjugate in kidney<br />

12 h after intravenous injection in mice.<br />

Arginine conjugation to the periphery of PAMAM dendrimer was reported to greatly increase<br />

the transfection efficiency. 74 Arginine residues of HIV-Tat peptide are thought to perform an<br />

important role in the intracellular translocation of molecular delivery. 75 The arginine-conjugated<br />

PAMAM dendrimer (PAMAM-Arg) showed enhanced gene expression in HepG2 cells, Neuro 2A<br />

cells, and primary rat vascular smooth muscle cells in comparison with native PAMAM dendrimer.<br />

It was presumed that the increased gene expression might be attributed to the difference in either the<br />

cell-penetrating activity during uptake, the nuclear localization efficiency after entry into the<br />

cytosol of arginine residues, or the synchronous function of both effects. Recently, Kang et al.<br />

conjugated Tat peptide to PAMAM dendrimer G5 for cellular delivery of antisense and siRNA<br />

oligonucleotides designed to inhibit MDR1 gene expression in vitro. 76 However, the dendrimer<br />

conjugate/oligonucleotide complexes weakly inhibited the gene expression, showing that conjugation<br />

of the dendrimer with the Tat cell-penetrating peptide failed to further enhance the<br />

effectiveness of the dendrimer.<br />

Lee et al. synthesized internally quaternized PAMAM-OH dendrimer (QPAMAM-OH) <strong>by</strong><br />

methylation of internal tertiary amines and investigated the gene delivery potency. 77 Although<br />

the dendrimer did not show high transfection efficiency in in vitro experiments, it could form a<br />

complex with plasmid DNA unlike with unmodified PAMAM-OH dendrimer, and its cytotoxicty<br />

was revealed to be very low in comparison with amine-terminated PAMAM dendrimer.<br />

PAMAM dendrimers were also applied for ex vivo or in vivo gene delivery. Hudde et al.<br />

investigated the efficiency of activated polyamidoamine dendrimers (SuperFecte) to transfect<br />

rabbit and human corneas in ex vivo culture. 78 Whole-thickness rabbit or human corneas were<br />

transfected ex vivo with complexes of dendrimers and plasmids containing lacZ or TNFR-Ig genes.<br />

Six to ten percent of the corneal endothelial cells expressed the marker gene, and corneas transfected<br />

with TNFR-Ig plasmid were found to express TNFR-Ig protein. The fifth generation of intact<br />

PAMAM dendrimers was also investigated for its ability to enhance gene transfer and expression in<br />

a clinically relevant murine vascularized heart transplantation model. 79 The complexes of the<br />

plasmid pMP6A-beta-gal, encoding beta-galactosidase (beta-Gal), and dendrimer were perfused<br />

via the coronary arteries during donor graft harvesting. The grafts infused with pMP6A-beta-gal/<br />

dendrimer complexes showed beta-Gal expression in myocytes from 7 to 14 days. The complexes<br />

containing 20 mg of DNA and 260 mg of dendrimer (1:20 charge ratio) in a total volume of 200 mL<br />

resulted in the highest gene expression in the grafts. Prolonged incubation (cold ischemic time) up<br />

to 2 h and pretreatment with serotonin were found to further enhance gene expression.<br />

Kukowska-Latallo et al. evaluated in vivo gene transfer and expression <strong>by</strong> intravascular and<br />

endobronchial routes using DNA complexed with G9 PAMAM dendrimer. 80 After intravenous<br />

administration of the dendrimer/pCF1CAT plasmid complexes, CAT expression was never<br />

observed in organs other than the lung. Repeated intravascular doses, administered four times at<br />

4-day intervals, maintained expression at 15–25% of peak concentrations achieved after the initial<br />

dose. However, in contrast with vascular delivery, intranasal delivery of the complex was found to<br />

give lower levels of CAT expression than that was observed with naked plasmid DNA. In situ<br />

localization of CAT enzymatic activity suggested that vascular administration seemed to achieve<br />

expression in the lung parenchyma, mainly within the alveoli, whereas endobronchial<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 501<br />

administration primarily targeted bronchial epithelium, showing that intravenously administered<br />

G9 dendrimer was an effective vector for pulmonary gene transfer and that transgene expression<br />

could be prolonged <strong>by</strong> repeated administration. PAMAM G4 dendrimer was also investigated for<br />

in vivo tumor targeting and transfection. 81 Biotinylated and 111 In labeled 20-mer antisense multiamino-linked<br />

oligonucleotide–avidin conjugate ( 111 In-oligo–Av) and the complexes of the<br />

oligonucleotide with PAMAM G4 ( 111 In-oligo/G4), biotinylated PAMAM G4–avidin<br />

( 111 In-oligo/G4–Av) were found to significantly enhance the tumor delivery of 111 In-oligo<br />

compared with delivery without carrier after i.p. injection.<br />

25.3.3 POLYPROPYLENIMINE DENDRIMERS<br />

Polypropylenimine (PPI) dendrimer is also a commercially available cationic dendrimer, but<br />

much less interest has been taken in its application to gene delivery systems compared to<br />

PAMAM dendrimer. Since Buhleier et al. synthesized low molecular weight PPI dendrimer in<br />

1978, 1 synthesis of the dendrimer has been improved through the Michael addition of amines to<br />

acrylonitrile, followed <strong>by</strong> the reduction of the nitriles to primary amines. 82 Diaminobutane (DAB)<br />

and diaminoethane cores are most commonly used for PPI dendrimer synthesis.<br />

The potential of PPI dendrimers as gene delivery carriers was first compared with other cationic<br />

polymers including branched PEI (25 kDa), linear PEI (22 kDa), and PAMAM dendrimer and<br />

others in Cos-7 cell line. 83 Although PPI dendrimer G5 (Astramole-64) showed very low transfection<br />

efficiency, MTT assay demonstrated that its cytotoxicity was relatively low compared with<br />

other polymers in treating polyplexes or free polycations. However, PPI dendrimers (generations 1–<br />

5) with DAB core were re-evaluated as efficient gene delivery systems <strong>by</strong> Uchegbu and coworkers.<br />

84 They reported that DNA binding increased with dendrimer generation through molecular<br />

modeling, and experimental data and cytotoxicity followed the trend DAB 64 (terminal amine<br />

number)ODAB 32ODAB 16ODOTAPODAB 4ODAB 8, whereas transfection efficacy<br />

followed the trend DAB 8ZDOTAPZDAB 16ODAB 4ODAB 32ZDAB 64. They concluded<br />

that the generation 2 polypropylenimine dendrimer (DAB 8) with a low level of cytoxicity gave<br />

optimum gene transfer activity.<br />

PPI dendrimers were also chemically modified for improved physiological properties. Park<br />

and coworkers investigated ternary complex of PPI dendrimer linked with DABs to its periphery<br />

(PPI–DAB), DNA, and cucurbituril (CB) for a non-covalent strategy in developing a gene<br />

delivery carrier. 85 CB is a large-cage compound composed of glycoluril units interconnected<br />

with methylene bridges, and it is able to form stable pseudorotaxanes with strings derived<br />

from DAB moeities of PPI–DAB through multiple non-covalent interactions. 86,87 The DNA<br />

binding capacities of PPI, PPI–DAB, and PPI–DAB/CB decreased with increasing generation<br />

and increased in the order, PPIOPPI–DABRPPI–DAB/CB. Fluorescence experiment using<br />

DAB-tethered acridine indicated that there was no change in the CB beads threaded on<br />

PPI–DAB after the addition of DNA to the PPI–DAB/CB binary complex. Importantly,<br />

in vitro transfection results showed that CB had no negative influence on the transfection efficiency<br />

of polymer/DNA binary complex, and the ternary complex was a highly active form of<br />

gene delivery vector, suggesting that this promising system can be applied to other non-covalent<br />

molecular recognitions, and it would be possible to create a really functional gene delivery carrier<br />

if the functionalized-CB were synthesized. Recently, another approach to developing gene<br />

delivery carriers with PPI dendrimers involved methyl quaternary ammonium derivatives of<br />

PPI dendrimers with varying generation. 88 Quaternization of DAB-cored PPI dendrimer (DAB<br />

8) with eight surface amines (Q 8) proved to be critical in enhancing DNA binding and resulted in<br />

improved colloidal stability and vector tolerability on i.v. injection of the polyplex.<br />

Quaternization was also found to improve in vitro cell biocompatiblity of DAB 16 and DAB 32<br />

polyplexes. Interestingly, the intravenous administration of DAB 16 and Q8 polyplexes resulted in<br />

liver-targeted gene expression as opposed to the lung-targeted gene expression obtained with the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


502<br />

control polymer, Exgen 500. This intrinsic targeting ability without the need for targeting ligands<br />

may be exploited for the targeted delivery of genes in the treatment of liver diseases. Also, Dufes<br />

et al. synthesized a vector system based on PPI dendrimer and investigated its transfection in tumors<br />

after i.v. injection. 89 The systemic tumor necrosis factor alpha (TNFalpha) gene delivery <strong>by</strong> the<br />

dendrimer was found to be efficient and non-toxic in the treatment of various established carcinomas.<br />

Interestingly, these dendrimers and other common polymeric transfection agents also<br />

exhibited plasmid-independent anti-tumor activity, ranging from pronounced growth retardation<br />

to complete tumor regression, indicating that the combination of the pharmacological active<br />

dendrimer and gene would be more potent.<br />

25.3.4 OTHER DENDRIMERS<br />

Nanotechnology for Cancer Therapy<br />

Dendritic poly(L-lysine)s (DPKs) of several generations were synthesized and investigated as gene<br />

transfection reagents. 90 The DPKs of the fifth and sixth generation showed efficient and serumtolerant<br />

gene transfection ability into several cell lines without significant cytotoxity. Several<br />

uptake inhibitor experiments suggested that the uptake of the DNA complex of DPK is mediated<br />

<strong>by</strong> the endocytosis pathway, especially the macropinocytotic process. In addition, terminal amino<br />

acids of DPK G6 were replaced <strong>by</strong> arginines and histidines to investigate the effect of substituting<br />

terminal cationic groups on the gene delivery. 91 Arginine-conjugated polylysine dendrimer<br />

(KGR6) showed higher transfection efficiency than unmodified KG6. In contrast, histidine-conjugated<br />

polylysine dendrimer (KGH6) showed high transfection efficiency only after polyplex<br />

formation under acidic conditions because of imdazole groups with a low pKa value of histidine<br />

residues. pH-dependent complex formation and transfection of KGH6 showed the potential for a<br />

functional gene delivery system.<br />

Choi et al. synthesized PEG–dendrimer conjugates PEG–poly(L-lysine) dendrimers and poly<br />

(L-lysine) dendrimer–PEG–poly(L-lysine) dendrimers (PLLD–PEG–PLLD) and characterized<br />

their self-assembly with plasmid DNA. 92,93 PEG was used as a core and poly(L-lysine) dendrimers<br />

were extended via a divergent peptide synthesis method. The dendrimers could form nanosized<br />

and water-soluble complexes with plasmid DNA in a generation-dependent manner and showed<br />

low cytotoxicity, but their transfection was not investigated. Thereafter, Kim et al. prepared<br />

triblock copolymers and PAMAM–PEG–PAMAM dendrimers for gene delivery (Figure 25.7). 94<br />

The copolymers showed the generation-dependent formation of nanosized complexes with<br />

plasmid DNA as PLLD–PEG–PLLD and high water-solubility because of their PEG cores. The<br />

G5 copolymer was found to possess low cytotoxicity and a much enhanced transfection ability in<br />

comparison to PLLD–PEG–PLLD in HepG2 cells and 293 cells.<br />

Joester et al. developed amphiphilic dendrimers with rigid tolane cores for gene delivery<br />

(Figure 25.8). 95 The dendrimers showed hydrophobicity, nominal charge-dependent DNA<br />

binding affinity, low cytotocity, and very high transfection efficiency superior to commercial<br />

transfection agents. However, the significant reduction of transfection efficiency with serum is<br />

thought to need research into the interaction of the dendrimer with anionic serum proteins.<br />

A series of water-soluble dendrimers with a phosphoramidothioate backbone (P-dendrimers)<br />

were prepared for gene delivery. 96 The polymers showed rather moderate cytotoxicity toward<br />

HeLa, HEK 293, and HUVEC cells, and they efficiently delivered oligodeoxyribonucleotide into<br />

HeLa cells in serum-containing medium.<br />

Hussain et al. developed another gene delivery system based on anionic dendrimer <strong>by</strong> directly<br />

conjugating ODNs to pentaerythritol-based phosphoroamidite dendrimer. 97 The cellular uptake<br />

of the ODN–dendrimer conjugates was up to 4-fold greater than for naked ODN in cancer cells.<br />

The ODN–dendrimer conjugates were found to effectively inhibit cancer cell growth, showing a<br />

marked knockdown in EGFR protein expression.<br />

Recently, Nishiyama et al. reported the application of photochemical internalization (PCI)mediated<br />

gene delivery in vivo using dendrimeric photosentisizer. 98 PCI is that the cytoplasmic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 503<br />

H2N NH2 NH H<br />

2<br />

2N NH<br />

H NH<br />

2N<br />

NH<br />

2<br />

NH<br />

NH NH<br />

2<br />

2<br />

NH<br />

NH<br />

H NH<br />

H NH<br />

2N N<br />

2<br />

N<br />

NH<br />

H NH<br />

2N HN<br />

NH<br />

2<br />

H N<br />

HN<br />

NH<br />

NH<br />

N<br />

H N<br />

N<br />

N<br />

H HN<br />

NH<br />

NH<br />

2N<br />

2<br />

N<br />

H NH<br />

2N 2<br />

HN<br />

NH<br />

N<br />

H N<br />

N<br />

HN N<br />

NH<br />

H<br />

H NH<br />

2N N<br />

NH<br />

2<br />

HN<br />

NH<br />

HN<br />

NH<br />

N<br />

N<br />

H NH2 2N N<br />

H<br />

NH<br />

NH<br />

HN<br />

HN<br />

NH<br />

N<br />

NH NH<br />

2 N<br />

2<br />

NH<br />

H NH2 2N HN<br />

NH<br />

NH<br />

NH<br />

NH H<br />

2<br />

2N H NH NH<br />

2N H2N 2<br />

2<br />

HN<br />

H2N NH2 HN<br />

NH HN<br />

HN O O<br />

O O<br />

O<br />

O HN<br />

N<br />

N<br />

O<br />

O<br />

N<br />

N<br />

O<br />

O<br />

HN<br />

HN<br />

O<br />

O<br />

O<br />

O<br />

N<br />

N<br />

N<br />

O<br />

O<br />

O<br />

HN<br />

O<br />

O<br />

O<br />

O<br />

N<br />

O<br />

HN<br />

N<br />

O O<br />

N O O N H<br />

O<br />

N<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O N<br />

O<br />

O<br />

O<br />

O<br />

O<br />

N<br />

O O<br />

N O O N H<br />

N<br />

HN<br />

O<br />

N H O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

N<br />

O<br />

O<br />

O<br />

N<br />

O<br />

O<br />

N<br />

HN<br />

H<br />

O<br />

O<br />

N<br />

O<br />

O<br />

N<br />

O HN<br />

O<br />

O O<br />

O<br />

O<br />

HN<br />

HN<br />

H<br />

H<br />

N<br />

FIGURE 25.7 Molecular structure of PAMAM–PEG–PAMAM G4.<br />

delivery of macromolecular compounds is enhanced <strong>by</strong> the photochemical disruption of the endosomal<br />

membrane in using light and a hydrophilic photosensitizer. 99 They prepared ternary<br />

complexes having cores of pDNA/cationic peptide polyplexes enveloped with anionic dendrimer<br />

phthalocyanine (DPc). The complexes showed greatly enhanced transfection efficiency and low<br />

H<br />

N<br />

O<br />

O<br />

HN<br />

NH<br />

FIGURE 25.8 Molecular structure of amphiphilic dendrimer with tolane core.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

N<br />

H<br />

O<br />

O<br />

N<br />

H<br />

O<br />

+ H3 N NH3 +<br />

HN<br />

O<br />

H<br />

N<br />

O<br />

NH<br />

NH 3 +<br />

NH 3 +<br />

NH 3 +<br />

NH 3 +<br />

NH 3 +<br />

NH 3 +<br />

NH 3 +


504<br />

photocytotoxicity in HeLa cells after photoirradiation, and they achieved significant gene<br />

expression in conjunctival tissue of rat eyes in contrast to ExGen500 that is one of the most<br />

efficient vectors.<br />

25.4 CONCLUSION<br />

In the past decade, various dendrimers have been developed and investigated for drug and gene<br />

delivery systems because of their unique structural properties such as multivalent and controllable<br />

functionalities on the periphery and highly branched and globular architectures as shown in this<br />

review. It is apparent that the application of dendrimers to drug and gene delivery systems has<br />

achieved considerably successful results, and much of the work reported in medical and pharmaceutical<br />

journals has demonstrated the growing interest in dendrimers.<br />

Although it is not presented in this chapter, the application of dendrimers has been extended in<br />

other medical fields such as diagnostics, antiviral, or antibacterial therapy. PAMAM dendrimerbased<br />

Gd III chelates were synthesized for use as a MRI contrast agent. 100,101 MRI is a powerful<br />

technique in medical diagnostics and used to visualize organs and blood vessels. The dendrimer<br />

chelates showed excellent MRI images and long blood circulation times in in vivo experiments.<br />

Also, chemically modified polylysine dendrimers and other dendrimers were synthesized for useful<br />

antiviral or antibacterial drugs. 102–105 In addition, numerous dendrimeric glycans or glycosylated<br />

dendrimers were shown to efficiently interact with natural carbohydrates receptors and were<br />

focused in their potential as microbial anti-adhesins, microbial toxin antagonists, and anti-cancer<br />

drugs. 106,107<br />

However, research about stable and safe biodistribution and efficient tissue localization as well<br />

as the biological function and fate of dendrimers is not well-established. Therefore, it is thought to<br />

be very important to explore the physiological safety and pharmacokinetics of developed dendrimers<br />

over a vaster range. It is also expected that a more convenient methodology to synthesizing<br />

novel dendrimers giving high water-solubility, good biocompatibility, and perfect structure will be<br />

designed for efficient application of dendrimers to biological systems.<br />

ACKNOWLEDGMENTS<br />

This work was supported <strong>by</strong> the National R&D Program Grant of The Ministry of Commerce,<br />

Industry and Energy (M10422010006-04N2201-00610), and the Gene Therapy Project of The<br />

Ministry of Science and Technology (M1053403004-05N3403-00410). The authors also wish to<br />

acknowledge helpful assistance received from Jung-un Baek.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Buhleier, E., Wehner, W., and Vögtle, F., Cascade- and nonskid-chain-like” syntheses of molecular<br />

cavity topologies, Synthesis, 155–158, 1978.<br />

2. Zeng, F. and Zimmerman, S. C., Dendrimers in supramolecular chemistry: From molecular recognition<br />

to self-assembly, Chem. Rev., 97, 1681–1712, 1977.<br />

3. Fischer, M. and Vögtle, F., Dendrimers: From design to application—a progress report, Angew.<br />

Chem. Int. Ed., 38, 884–905, 1999.<br />

4. Stiriba, S. E., Frey, H., and Haag, R., Dendritic polymers in biomedical applications: From potential<br />

to clinical use in diagnostics and therapy, Angew. Chem. Int. Ed., 41, 1329–1334, 2002.<br />

5. Patri, A. K., Majoros, I. J., and Baker, J. R. Jr., Dendritic polymer macromolecular carriers for drug<br />

delivery, Curr. Opin. Chem. Biol., 6, 466–471, 2002.<br />

6. Gilles, E. R. and Frechet, J. M. J., Dendrimers and dendritic polymers in drug delivery, Drug Discov.<br />

Today, 10, 35–43, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 505<br />

7. Blume, G. and Cevc, G., Liposomes for sustained drug release in vivo, Biochim. Biophys. Acta, 1029,<br />

97, 1990.<br />

8. Lian, T. and Ho, R. J. Y., Trends and developments in liposome drug delivery systems, J. Pharm.<br />

Sci., 90, 680, 2001.<br />

9. Rosler, A., Vandermeulen, G. W. M., and Klok, H. A., Advanced drug delivery devices via selfassembly<br />

of amphiphilic block copolymers, Adv. Drug Deliv. Rev., 53, 95–108, 2001.<br />

10. Newkome, G. R. et al., Unimolecular micelles, Angew. Chem. Int. Ed. Engl., 30, 1178–1180, 1991.<br />

11. Jansen, J. F. G. A., de Brabander-van den Berg, E. M. M., and Meijer, E. W., Encapsulation of guest<br />

molecules into a dendritic box, Science, 266, 1226–1229, 1994.<br />

12. Jansen, J. F. G. A., Meijer, E. W., and de Brabander-van den Berg, E. M. M., The dendritic box:<br />

Shape-selective liberation of encapsulated guests, J. Am. Chem. Soc., 117, 4417–4418, 1995.<br />

13. Baars, M. W. P. L. et al., The localization of guests in water-soluble oligoethyleneoxy-modified<br />

poly(propylene imine) dendrimers, Angew. Chem. Int. Ed., 39, 1285–1288, 2000.<br />

14. Sideratou, Z., Tsiourvas, D., and Paleos, C. M., Solubilization and release properties of PEGylated<br />

diaminobutane poly(propylene imine) dendrimers, J. Colloid Interface Sci., 242, 272–276, 2001.<br />

15. Sideratou, A., Tsiourvas, D., and Paleos, C. M., Quaternized poly(propylene imine) dendrimers as<br />

novel pH-sensitive controlled-release systems, Langmuir, 16, 1766–1769, 2000.<br />

16. Tomalia, D. A. et al., A new class of polymers: Starbust-dendritic macromolecules, Polym. J., 17,<br />

117–132, 1985.<br />

17. Twyman, L. J. et al., The synthesis of water soluble dendrimers, and their application as possible drug<br />

delivery systems, Tetrahedron Lett., 40, 1743–1746, 1999.<br />

18. Kojima, C. et al., Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and<br />

their ability to encapsulate anticancer drugs, Bioconjug. Chem., 11, 910–917, 2000.<br />

19. Bhadra, D. et al., A PEGylated dendritic nanoparticle carrier of fluorouracil, Int. J. Pharm., 257,<br />

111–124, 2003.<br />

20. Haba, Y. et al., Synthesis of biocompatible dendrimers with a peripheral network formed <strong>by</strong> linking<br />

of polymerizable groups, Polymer, 46, 1813–1820, 2005.<br />

21. Liu, M., Kono, K., and Frechet, J. M. J., Water-soluble unimolecular micelles: Their potential as<br />

drug delivery agents, J. Control. Release, 65, 121–131, 2000.<br />

22. Gillies, E. R., Jonsson, T. B., and Frechet, J. M. J., Stimuli-responsive supramolecular assemblies of<br />

linear-dendritic copolymers, J. Am. Chem. Soc., 126, 11936–11943, 2004.<br />

23. Namazi, H. and Adeli, M., Dendrimers of citric acid and poly(ethylene glycol) as the new drugdelivery<br />

agents, Biomaterials, 26, 1175–1183, 2005.<br />

24. Roberts, J. C. et al., Using starbust dendrimers as linker molecules to radiolabel antibodies,<br />

Bioconjug. Chem., 1, 305–308, 1990.<br />

25. Wu, C. et al., Metal–chelate–dendrimer–antibody constructs for use in radioimmunotherapy and<br />

imaging, Bioorg. Med. Chem. Lett., 4, 449–454, 1994.<br />

26. Patri, A. K. et al., Antibody–dendrimer conjugates for targeted prostate cancer therapy, Polym.<br />

Mater. Sci. Eng., 86, 130, 2002.<br />

27. Duncan, R. and Malik, N., Dendrimers: Biocompatibility and potential for delivery of anticancer<br />

agents, Proc. Int. Symp. Control. Release Bioact. Mater., 23, 105–106, 1996.<br />

28. Malik, N., Evagorou, E. G., and Duncan, R., Dendrimer–platinate: A novel approach to cancer<br />

chemotherapy, Anticancer Drugs, 10, 767–776, 1999.<br />

29. Zhuo, R. X., Du, B., and Lu, Z. R., In vitro release of 5-fluorouracil with cyclic core dendritic<br />

polymer, J. Control. Release, 57, 249–257, 1999.<br />

30. Wang, D. et al., Synthesis of starlike N-(2-hydroxypropyl)methacrylamide copolymers: Potential<br />

drug carriers, Biomacromolecules, 1, 313–319, 2000.<br />

31. Wiwattanapatapee, R., Lomlim, L., and Saramunee, K., Dendrimers conjugates for colonic delivery<br />

of 5-aminosalicylic acid, J. Control. Release, 88, 1–9, 2003.<br />

32. Quintana, A. et al., Design and function of a dendrimer-based therapeutic nanodevice targeted to<br />

tumor cells through the folate receptor, Pharm. Res., 19, 1310–1316, 2002.<br />

33. Tansey, W. et al., Synthesis and characterization of branched poly(L-glutamic acid) as a biodegradable<br />

drug carrier, J. Control. Release, 94, 39–51, 2004.<br />

34. Kono, K., Liu, M., and Fréchet, J. M. J., Design of dendritic macromolecules containing folate or<br />

methotrexate residues, Bioconjug. Chem., 10, 1115–1121, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


506<br />

Nanotechnology for Cancer Therapy<br />

35. Liu, M., Kono, K., and Fréchet, J. M. J., Water-soluble dendrimer–polyethylene glycol starlike<br />

conjugates as potential drug carriers, J. Polym. Sci. A, 37, 3492–3503, 1999.<br />

36. Choe, Y. H. et al., Anticancer drug delivery systems: Multi-loaded N 4 -acyl poly(ethylene glycol)<br />

prodrugs of ara-C. II. Efficacy in ascites and solid tumors, J. Control. Release, 79, 55–70, 2002.<br />

37. Schiavon, O. et al., PEG–ara-C conjugates for controlled release, Eur. J. Med. Chem., 39, 123–133, 2004.<br />

38. Padilla De Jesu’s, O. L. et al., Polyester dendritic systems for drug delivery applications: In vitro and<br />

in vivo evaluation, Bioconjug. Chem., 13, 453–461, 2002.<br />

39. de Groot, F. M. H. et al., “Cascade-release dendrimers” liberation all end groups upon a single<br />

triggering event in the dendritic core, Angew. Chem. Int. Ed., 42, 4490–4494, 2003.<br />

40. Amir, R. J. et al., Self-immolative dendrimers, Angew. Chem. Int. Ed., 42, 4494–4499, 2003.<br />

41. Shamis, M., Lode, H. N., and Shabat, D., Bioactivation of self-immolative dendritic prodrugs <strong>by</strong><br />

catalytic antibody 38C2, J. Am. Chem. Soc., 126, 1726–1731, 2004.<br />

42. Hawthorne, M. F., The role of chemistry in the development of boron neutron capture therapy of<br />

cancer, Angew. Chem. Int. Ed. Engl., 32, 950–984, 1993.<br />

43. Barth, R. F. et al., Boronated starbust dendrimer–monoclonal antibody immunoconjugates: Evaluation<br />

as a potential delivery system for neutron capture therapy, Bioconjug. Chem., 5, 58–66, 1994.<br />

44. Capala, J. et al., Boronated epidermal growth factor as a potential targeting agent for boron neutron<br />

capture therapy of brain tumors, Bioconjug. Chem., 7, 7–15, 1996.<br />

45. Yang, W. et al., Intratumoral delivery of boronated epidermal growth factor for neutron capture<br />

therapy of brain tumors, Cancer Res., 57, 4333–4339, 1997.<br />

46. Wu, G. et al., Site-specific conjugation of boron-containing dendrimers to anti-EGF receptor monoclonal<br />

antibody cetuximab(IMC-C225) and its evaluation as a potential delivery agent for neutron<br />

capture therapy, Bioconjug. Chem., 15, 185–194, 2004.<br />

47. Shukla, S. et al., Synthesis and biological evaluation of folate receptor-targeted boronated PAMAM<br />

dendrimers as potential agents for neutron capture therapy, Bioconjug. Chem., 14, 158–167, 2003.<br />

48. Qualmann, B. et al., Synthesis of boron-rich lysine dendrimers as protein labels in electron<br />

microscopy, Angew. Chem. Int. Ed. Engl., 35, 909–911, 1996.<br />

49. Matsumoto, T., Transport calculations of depth–dose distributions for gadolinium neutron capture<br />

therapy, Phys. Med. Biol., 37, 155–162, 1992.<br />

50. Khokhlov, V. F. et al., Neutron capture therapy with gadopentetate dimeglumine: Experiments on<br />

tumor-bearing rats, Acad. Radiol., 2, 392–398, 1995.<br />

51. Hofmann, B. et al., Gadolinium neutron capture therapy (GdNCT) of melanoma cells and solid<br />

tumors with the magnetic resonance imaging contrast agent Gadobutrol, Invest. Radiol., 34,<br />

126–133, 1999.<br />

52. Caravan, P. et al., Gadolinium(III) chelates as MRI contrast agents: Structure, dynamics, and applications,<br />

Chem. Rev., 99, 2293–2352, 1999.<br />

53. Kobayashi, H. et al., Avidin-dendrimer-(1B4M-Gd)254: A tumor-targeting therapeutic agent for<br />

gadolinium neutron capture therapy of intraperitoneal disseminated tumor which can be monitored<br />

<strong>by</strong> MRI, Bioconjug. Chem., 12, 587–593, 2001.<br />

54. Dougherty, T. J. et al., Photodynamic therapy, J. Natl. Cancer Inst., 90, 889–905, 1998.<br />

55. Battah, S. H. et al., Synthesis and biological studies of 5-aminolevulinic acid-containing dendrimers<br />

for photodynamic therapy, Bioconjug. Chem., 12, 980–988, 2001.<br />

56. Peng, Q. et al., 5-Aminolevulinic acid-based photodynamic therapy: Principles and experimental<br />

research, Photochem. Photobiol., 65, 235–251, 1997.<br />

57. Nishiyama, N. et al., Light-harvesting ionic dendrimer porphyrins as new photosensitizers for photodynamic<br />

therapy, Bioconjug. Chem., 14, 58–66, 2003.<br />

58. Zhang, G. D. et al., Polyion complex micelles entrapping cationic dendrimer porphyrin: Effective<br />

photosensitizer for photodynamic therapy of cancer, J. Control. Release, 93, 141–150, 2003.<br />

59. Dichtel, W. R. et al., Singlet oxygen generation via two-photon excited FRET, J. Am. Chem. Soc.,<br />

126, 5380–5381, 2004.<br />

60. Boussif, O. et al., A versatile vector for gene and oligonucleotide transfer into cells in culture and<br />

in vivo: Polyethylenimine, Proc. Natl Acad. Sci., 92, 7297–7301, 1995.<br />

61. Rejman, J., Bragonzi, A., and Conese, M., Role of clathrin- and caveolae-mediated endocytosis in<br />

gene transfer mediated <strong>by</strong> lipo- and polyplexes, Mol. Ther., 12, 468–474, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimers as Drug and Gene Delivery Systems 507<br />

62. Haensler, J. and Szoka, F. C. Jr., Polyamidoamine cascade polymers mediate efficient transfection of<br />

cells in culture, Bioconjug. Chem., 4, 372–379, 1993.<br />

63. Kukowska-Latallo, J. F. et al., Efficient transfer of genetic material into mammalian cells using<br />

starburst polyamidoamine dendrimers, Proc. Natl Acad. Sci. USA, 93, 4897–4902, 1996.<br />

64. Harada, Y. et al., Highly efficient suicide gene expression in hepatocellular carcinoma cells <strong>by</strong><br />

Epstein–Barr virus-based plasmid vectors combined with polyamidoamine dendrimer, Cancer<br />

Gene Ther., 7, 27–36, 2000.<br />

65. Tanaka, S. et al., Targeted killing of carcinoembryonic antigen (CEA)-producing cholangiocarcinoma<br />

cells <strong>by</strong> polyamidoamine dendrimer-mediated transfer of an Epstein–Barr virus (EBV)-based<br />

plasmid vector carrying the CEA promoter, Cancer Gene Ther., 7, 1241–1250, 2000.<br />

66. Tang, M. X., Redemann, C. T., and Szoka, F. C. Jr., In vitro gene delivery <strong>by</strong> degraded polyamidoamine<br />

dendrimers, Bioconjug. Chem., 7, 703–714, 1996.<br />

67. Bielinska, A. et al., Regulation of in vitro gene expression using antisense oligonucleotides or<br />

antisense expression plasmids transfected using starburst PAMAM dendrimers, Nucleic Acids<br />

Res., 24, 2176–2182, 1996.<br />

68. Delong, R. et al., J. Pharm. Sci., 86, 762–764, 1997.<br />

69. Yoo, H. and Juliano, R. L., Enhanced delivery of antisense oligonucleotides with fluorophoreconjugated<br />

PAMAM dendrimers, Nucleic Acids Res., 28, 4225–4231, 2000.<br />

70. Maksimenko, A. V. et al., Optimisation of dendrimer-mediated gene transfer <strong>by</strong> anionic oligomers,<br />

J. Gene Med., 5, 61–71, 2003.<br />

71. Roessler, B. J. et al., Substituted b-cyclodextrins interact with PAMAM dendrimer–DNA complexes<br />

and modify transfection efficiency, Biochem. Biophys. Res. Commun., 283, 124–129, 2001.<br />

72. Kihara, F. et al., Effects of structure of polyamidoamine dendrimer on gene transfer efficiency of the<br />

dendrimer conjugate with a-cyclodextrin, Bioconjug. Chem., 13, 1211–1214, 2002.<br />

73. Wada, K. et al., Improvement of gene delivery mediated <strong>by</strong> mannosylated dendrimer/a-cyclodextrin<br />

conjugates, J. Control. Release, 104, 397–413, 2005.<br />

74. Choi, J. S. et al., Enhanced transfection efficiency of PAMAM dendrimer <strong>by</strong> surface modification<br />

with L-arginine, J. Control. Release, 99, 445–456, 2004.<br />

75. Tung, C. H. and Weissleder, R., Arginine containing peptides as delivery vectors, Adv. Drug Deliv.<br />

Rev., 55, 281–294, 2003.<br />

76. Kang, H. et al., Tat-conjugated PAMAM dendrimers as delivery agents for antisense and siRNA<br />

oligonucleotides, Pharm. Res., 22, 2099–2106, 2005.<br />

77. Lee, J. H. et al., Polyplexes assembled with internally quaternized PAMAM-OH dendrimer and<br />

plasmid DNA have a neutral surface and gene delivery potency, Bioconjug. Chem., 14,<br />

1214–1221, 2003.<br />

78. Hudde, T. et al., Activated polyamidoamine dendrimers, a non-viral vector for gene transfer to the<br />

corneal endothelium, Gene Ther., 6, 939–943, 1999.<br />

79. Wang, Y. et al., DNA/dendrimer complexes mediate gene transfer into murine cardiac transplants ex<br />

vivo, Mol. Ther., 2, 602–608, 2000.<br />

80. Kukowska-Latallo, J. F. et al., Intravascular and endobronchial DNA delivery to murine lung tissue<br />

using a novel, nonviral vector, Hum. Gene Ther., 11, 1385–1395, 2000.<br />

81. Sato, N. et al., Tumor targeting and imaging of intraperitoneal tumors <strong>by</strong> use of antisense oligo-DNA<br />

complexed with dendrimers and/or avidin in mice, Clin. Cancer Res., 7, 3606–3612, 2001.<br />

82. de Brabander-van den Berg, E. M. M. and Meijer, E. W., Poly(propylene imine) dendrimers: Largescale<br />

synthesis <strong>by</strong> hetereogeneously catalyzed hydrogenations, Angew. Chem. Int. Ed. Engl., 32,<br />

1308–1311, 1993.<br />

83. Gebhart, C. L. and Kabanov, A. V., Evaluation of polyplexes as gene transfer agents, J. Control.<br />

Release, 73, 401–416, 2001.<br />

84. Zinselmeyer, B. H. et al., The lower-generation polypropylenimine dendrimers are effective genetransfer<br />

agents, Pharm. Res., 19, 960–967, 2002.<br />

85. Lim, Y. b. et al., Self-assembled ternary complex of cationic dendrimer, cucurbituril, and<br />

DNA: Noncovalent strategy in developing a gene delivery carrier, Bioconjug. Chem., 13,<br />

1181–1185, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


508<br />

Nanotechnology for Cancer Therapy<br />

86. Whang, D. et al., Self-assembly of a polyrotaxane containing a cyclic “bead” in every structural unit<br />

in the solid state: Cucurbituril molecules threaded on a one-dimensional coordination polymer,<br />

J. Am. Chem. Soc., 118, 11333–11334, 1996.<br />

87. Lee, J. W. et al., Novel pseudorotaxane-terminated dendrimers: Supramolecular modification of<br />

dendrimer periphery, Angew. Chem. Int. Ed., 40, 746–749, 2001.<br />

88. Schatzlein, A. G. et al., Preferential liver gene expression with polypropylenimine dendrimers,<br />

J. Control. Release, 101, 247–258, 2005.<br />

89. Dufes, C. et al., Synthetic anticancer gene medicine exploits intrinsic antitumor activity of cationic<br />

vector tocure established tumors, Cancer Res., 65, 8079–8084, 2005.<br />

90. Ohsaki, M. et al., In vitro gene transfection using dendritic poly(L-lysine), Bioconjug. Chem., 13,<br />

510–517, 2002.<br />

91. Okuda, T. et al., Characters of dendritic poly(L-lysine) analogues with the terminal lysines replaced<br />

with arginines and histidines as gene carriers in vitro, Biomaterials, 25, 537–544, 2004.<br />

92. Choi, J. S. et al., Poly(ethylene glycol)-block-poly(L-lysine) dendrimer: Novel linear polymer/dendrimer<br />

block copolymer forming a spherical water-soluble polyionic complex with DNA, Bioconjug.<br />

Chem., 10, 62–65, 1999.<br />

93. Choi, J. S. et al., Synthesis of a barbell-like triblock copolymer, poly(L-lysine) dendrimer-blockpoly(ethylene<br />

glycol)-block-poly(L-lysine) dendrimer, and its self-assembly with plasmid DNA,<br />

J. Am. Chem. Soc., 122, 474–480, 2000.<br />

94. Kim, T. I. et al., PAMAM–PEG–PAMAM: Novel triblock copolymer as a biocompatible and<br />

efficient gene delivery carrier, Biomacromolecules, 5, 2487–2492, 2004.<br />

95. Joester, D. et al., Amphiphilic dendrimers: Novel self-assembling vectors for efficient gene delivery,<br />

Angew. Chem. Int. Ed., 42, 1486–1490, 2003.<br />

96. Maszewska, M. et al., Water-soluble polycationic dendrimers with a phosphoramidothioate backbone:<br />

Preliminary studies of cytotoxicity and oligonucleotide/plasmid delivery in human cell<br />

culture, Oligonuleotides, 13, 193–205, 2003.<br />

97. Hussain, M. et al., A novel anionic dendrimer for improved cellular delivery of antisense oligonucleotides,<br />

J. Control. Release, 99, 139–155, 2004.<br />

98. Nishiyama, N. et al., Light-induced gene transfer from packaged DNA enveloped in a dendrimeric<br />

photosensitizer, Nat. Mater., 4, 934–941, 2005.<br />

99. Høgest, A. et al., Photochemical internalization in drug and gene delivery, Adv. Drug Deliv. Rev., 56,<br />

95–115, 2004.<br />

100. Wiener, E. C. et al., Dendrimer-based metal chelates: A new class of magnetic resonance imaging<br />

contrast agents, Magn. Reson. Med., 31, 1–8, 1994.<br />

101. Bryant, L. H. Jr. et al., Synthesis and relaxometry of high-generation (GZ5, 7, 9, and 10) PAMAM<br />

dendrimer–DOTA–gadolinium chelates, J. Magn. Reson. Imaging, 9, 348–352, 1999.<br />

102. Bourne, N. et al., Dendrimers, a new class of candidate topical microbicides with activity against<br />

herpes simplex virus infection, Antimicrob. Agents Chemother., 44, 2471–2474, 2000.<br />

103. Witvrouw, M. et al., Potent anti-HIV (type 1 and type 2) activity of polyoxometalates: Structure–<br />

activity relationship and mechanism of action, J. Med. Chem., 43, 778–783, 2000.<br />

104. Landers, J. J. et al., Prevention of influenza pneumonitis <strong>by</strong> sialic acid-conjugated dendritic polymers,<br />

J. Infect. Dis., 186, 1222–1230, 2002.<br />

105. Chen, C. Z. S. and Cooper, S. L., Recent advances in antimicrobial dendrimers, Adv. Mater., 12,<br />

843–846, 2000.<br />

106. Veprek, P. and Jezek, J., Peptide and glycopeptide dendrimers. Part II, J. Pept. Sci., 5, 203–220,<br />

1999.<br />

107. Andre, S. et al., Wedgelike glycodendrimers as inhibitors of binding of mammalian galectins to<br />

glycoproteins, lactose maxiclusters, and cell surface glycoconjugates, Chembiochem, 2, 822–830,<br />

2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


26<br />

CONTENTS<br />

Dendritic Nanostructures<br />

for Cancer Therapy<br />

Ashootosh V. Ambade, Elamprakash N. Savariar,<br />

and S. (Thai) Thayumanavan<br />

26.1 Introduction ....................................................................................................................... 509<br />

26.2 Dendrimer–Drug Conjugates ............................................................................................ 511<br />

26.3 Encapsulation of the Drug in a Dendrimer....................................................................... 514<br />

26.4 Polymeric Drug Carriers Based on Dendrimers ............................................................... 517<br />

26.5 Summary............................................................................................................................ 518<br />

References..................................................................................................................................... 518<br />

26.1 INTRODUCTION<br />

Nanomaterials have emerged from academic curiosity and are being applied in many diverse areas<br />

such as information technology and biotechnology. 1–4 Synthetic nanomaterials based on macromolecules<br />

are relevant in biomedical applications because their size scale is comparable with that of<br />

biomacromolecules. 5 Dendrimers provide a prominent example of such synthetic macromolecules<br />

that have attracted great attention in recent years because of their potential in drug delivery among<br />

other promising applications. 6–12 Dendrimers are highly branched polymers of a few nanometer<br />

dimensions that possess a globular shape at high molecular weights. 13,14 These are synthesized <strong>by</strong><br />

repetitive organic transformations and, as a result, are almost monodisperse. An interesting structural<br />

feature of dendrimers is the spatially isolated core in the dendritic interior that is reminiscent<br />

of site-isolation in biomolecules such as proteins. The branches emanating from this core provide<br />

cavities in the interior that can be exploited for guest encapsulation. Every layer of branches that is<br />

added to the core during the synthesis is termed as a generation. The highly branched nature of<br />

dendrimers leads to large number of peripheral groups. These surface functionalities can be<br />

modified <strong>by</strong> attachment of drugs or targeting ligands to meet specific applications that often<br />

involve controlled interaction with cell walls and other biological targets. A representative structure<br />

is shown in Figure 26.1 that illustrates the typical features—core, intermediate layers, and peripheral<br />

groups—of a dendrimer. The dendritic structure in terms of type of repeat units, size of interior<br />

cavities, and type of peripheral groups can be tuned <strong>by</strong> using either convergent or divergent<br />

synthetic strategy. 15,16 It is then obvious that such versatile molecules are being intensely<br />

pursued in cancer therapeutics. 17,18<br />

A major problem with anti-cancer agents is that they indiscriminately attack healthy cells as<br />

well as cancerous cells. These harmful side effects can be reduced <strong>by</strong> developing a drug delivery<br />

vehicle that is specific to tumor cells. This can be achieved <strong>by</strong> employing a strategy called active<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

509


510<br />

H 2N<br />

H 2N<br />

H<br />

N<br />

O<br />

H<br />

N<br />

Intermediate layers<br />

(dendritic interior)<br />

H 2N<br />

N<br />

H<br />

H 2N<br />

H 2N<br />

O<br />

O<br />

N<br />

HN<br />

N<br />

O<br />

O<br />

NH<br />

H<br />

N<br />

O<br />

NH<br />

H 2N<br />

O<br />

N<br />

O<br />

HN<br />

N<br />

N<br />

NH 2<br />

NH<br />

O<br />

NH<br />

O<br />

NH<br />

N<br />

O<br />

NH O<br />

HN O<br />

O<br />

HN<br />

NH 2<br />

N<br />

Core<br />

targeting wherein functionalities that respond to over-expressed receptors on tumor cells are<br />

attached to the drug carrier. Dendrimers provide an opportunity in this direction in the form of a<br />

large number of tunable peripheral groups where a variety of functionalities can be incorporated<br />

(Figure 26.2). Another strategy is the passive targeting that relies on the well-known enhanced<br />

permeation and retention (EPR) effect. This effect, first described <strong>by</strong> Maeda et al., 19 takes advantage<br />

of poorly formed (leaky) vasculature of solid tumors that allows accumulation of polymer–drug<br />

HN<br />

O<br />

N<br />

O<br />

HN<br />

O<br />

HN<br />

O<br />

NH<br />

O<br />

N<br />

N<br />

N<br />

O<br />

NH<br />

H 2N<br />

Peripheral amine groups<br />

N<br />

NH<br />

NH 2<br />

HN<br />

H<br />

N<br />

O<br />

O<br />

NH<br />

H 2N<br />

O<br />

O<br />

NH<br />

N<br />

O<br />

O<br />

N<br />

NH 2<br />

H<br />

N<br />

H<br />

N<br />

O<br />

O<br />

N<br />

H<br />

NH<br />

FIGURE 26.1 Structural features of a dendrimer are illustrated with a structure of amine-terminated PAMAM<br />

dendrimer.<br />

= Drug molecule<br />

= Targetingmoiety<br />

= Functionality for imaging<br />

FIGURE 26.2 A dendrimer with multiple functionalities attached to it is illustrated <strong>by</strong> a cartoon.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

NH 2<br />

NH 2<br />

NH 2<br />

NH 2


Dendritic Nanostructures for Cancer Therapy 511<br />

conjugates ranging in size between 10 and 500 nm within tumors than of free drugs. The<br />

polymeric molecules are retained following the accumulation due to their larger size whereas<br />

free drug molecules are easily eliminated from the cells. Dendrimers have an advantage for this<br />

strategy as well because drugs can be physically encapsulated into the interior of nanometer-sized<br />

dendrimers, there<strong>by</strong> increasing the uptake of drugs into tumors. Before describing the various ways<br />

in which dendritic structure has been utilized for drug delivery in cancer therapy, it is useful to<br />

discuss the techniques that are employed for this purpose.<br />

Photodynamic therapy (PDT), also known as photoradiation therapy or photochemotherapy,<br />

involves the use of a light-activated drug called photosensitizer that can be administered orally or<br />

intraperitoneally, or it can be applied to the skin. 20,21 When light of a specific wavelength that is<br />

absorbed <strong>by</strong> the dye is shined on the targeted tissue, the photosensitizer is excited from ground<br />

singlet state to excited singlet state. It then undergoes intersystem crossing to a longer-lived excited<br />

triplet state that reacts with the ground triplet state of molecular oxygen present in the tissues. This<br />

energy transfer creates highly-reactive excited singlet-state oxygen species. The singlet oxygen<br />

reacts destructively with near<strong>by</strong> biomolecules and induces cell death either <strong>by</strong> apoptosis or<br />

necrosis. PDT, however, has limitations because it is applicable only in areas where light can reach.<br />

Neutron capture therapy (NCT) is another promising technique for treatment of cancer. 22 It is<br />

based on the nuclear reaction that occurs when a stable isotope, for example, boron 10 B, is irradiated<br />

with low energy or thermal neutrons to produce high-energy alpha particles ( 4 He nuclei) and 7 Li<br />

ions. The combined path length of these two particles is about 12 mm that is close to the diameter of<br />

a cell. Therefore, if the tumor cells can be supplied with a high concentration of 10 B atoms, the<br />

effect of the radiation will be limited to tumor cells, and the normal cells may be spared. This is<br />

particularly important in cancer therapy where prevention of indiscriminate destruction of healthy<br />

cells is a challenge. 157 Gd is another isotope of interest because of its large neutron capture cross<br />

section. Success of the above two therapies depends on localizing sufficient concentration of the<br />

photosensitizer or radioactive material in the tumor cells while having negligible accumulation in<br />

the healthy cells.<br />

Magnetic resonance imaging (MRI) is one of the most widely used imaging techniques that has<br />

greatly improved medical diagnosis. 23 It is based on the relaxation properties of hydrogen nuclei in<br />

water in body tissues when placed in a uniform magnetic field. In clinical practice, for detection of<br />

cancer in particular, it uses the fact that the relaxation times of protons in cancerous and normal<br />

tissues markedly differ. One of the advantages of MRI over other imaging techniques is that it is a<br />

non-radiative technique and harmless to the patient.<br />

This chapter will focus on the structural aspects of dendritic molecules that have been used to<br />

design efficient and multifunctional drug delivery vehicles in cancer therapy with the help of<br />

selected illustrative examples. There are only few cases where these macromolecules have been<br />

studied in vivo, i.e., in animal models, and the biological studies will be discussed only cursorily.<br />

26.2 DENDRIMER–DRUG CONJUGATES<br />

The term dendrimer–drug conjugate is mainly applied when a drug molecule is covalently linked to<br />

a dendrimer. The drug can be covalently attached to a dendrimer on the periphery or at the core and<br />

quite rarely at the branching points, i.e., in the interior layers. Attachment of a drug to multiple<br />

peripheral groups in a dendrimer greatly increases the effective concentration of a drug once it is<br />

released at the targeted site. This is particularly advantageous for the use of prodrugs. In the case of<br />

PDT, however, it is desirable to isolate the drug molecule to prevent fluorescence quenching<br />

because of aggregation, and this can be achieved <strong>by</strong> attaching the photosensitizer to the core of<br />

a dendrimer. A dendrimer–drug conjugate can provide certain advantages over conventional polymeric<br />

drug delivery vehicles because dendrimers are monodisperse, structurally defined<br />

macromolecules with a controlled molecular weight and size.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


512<br />

O 2N<br />

O<br />

HN<br />

O<br />

O<br />

O<br />

O<br />

OR<br />

O<br />

HN O<br />

O<br />

O<br />

O<br />

OR<br />

OR<br />

O<br />

OR<br />

O<br />

R = Ph<br />

O OH<br />

When a drug is attached to peripheral groups of a dendrimer, the linker between drug and<br />

dendrimer is of crucial importance because the drugs need to be efficiently released in active form at<br />

the targeted site. Advantage is taken of the acidic or reducing environment near the tumor cells <strong>by</strong><br />

attaching the drugs to dendritic periphery via acid-labile or disulphide (susceptible to reduction)<br />

linker (see Section 26.4 for examples). In this case, each drug is separately cleaved over a long<br />

period of time. An elegant design was reported in this direction wherein a single triggering event at<br />

the core results in disassembly of the dendrimer and release of all the paclitaxel molecules attached<br />

to the periphery. 24 The drug release in these dendrimers incorporating a masked 4-aminocinnamyl<br />

alcohol linker is activated <strong>by</strong> reduction of nitro to amino group under mild reducing conditions<br />

(Figure 26.3).<br />

Poly(amidoamine) (PAMAM) dendrimers have been widely used for preparation of<br />

dendrimer–drug conjugates because of easily modifiable terminal amine groups, commercial availability<br />

of higher generation dendrimers, and biocompatibility. 25–27 For example,<br />

methylprednisolone (MP) was attached to PAMAM periphery through a glutaric acid linker. 28<br />

Because the MP–glutaric acid compound was attached to a preformed dendrimer, only 12 drug<br />

molecules (32 wt%) could be conjugated to a dendrimer that has 32 terminal moieties as a result of<br />

steric hindrance. However, considering that MP is a bulky steroidal drug, this was a significant<br />

increase over earlier payloads.<br />

Monoclonal antibodies (mAb) are useful for targeted delivery because these can specifically<br />

bind to certain receptors over-expressed on tumor cells. Antibody–PAMAM conjugates have been<br />

prepared <strong>by</strong> attaching J591 anti-PSMA antibody to a fluorescein-labeled dendrimer for targeted<br />

delivery in prostate cancer therapy. It was based on the fact that prostate-specific membrane antigen<br />

(PSMA) is over-expressed on tumor cells than on normal tissue. 29 Attachment of multiple functionalities<br />

for targeting (folic acid) and imaging (fluorescein) along with an anti-cancer drug<br />

(methotrexate) to a dendrimer periphery has also been demonstrated with PAMAM dendrimers 30<br />

and is elaborated below with examples on MRI and NCT. Similarly, different PAMAM dendrimers<br />

attached with fluorescein and folic acid were covalently attached to oligonucleotides and assembled<br />

via hybridization to afford multifunctional dendrimer clusters as illustrated <strong>by</strong> a cartoon in<br />

Figure 26.4. 31 On a different note, carbohydrate moieties have been attached to PAMAM dendrimers<br />

to study multivalent protein–carbohydrate interactions. This is particularly relevant to cancer<br />

metastasis that involves these interactions in cellular recognition processes. 32–34<br />

O<br />

N<br />

H<br />

Nanotechnology for Cancer Therapy<br />

Ph<br />

O<br />

O<br />

AcO<br />

H<br />

OH AcO<br />

OBz<br />

Paclitaxel<br />

Point of attachment to the dendrimer<br />

FIGURE 26.3 Paclitaxel–dendrimer conjugate with a cascade-release mechanism. The NO 2 group at the focal<br />

point is converted to NH2 under reducing conditions triggering the cascade-release.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O


Dendritic Nanostructures for Cancer Therapy 513<br />

FITC<br />

PAMAM dendrimer<br />

FITC<br />

11 nm<br />

20 nm<br />

FITC-functionality for imaging<br />

FA-functionality for targeting<br />

PAMAM dendrimers have also been used for preparation of MRI contrast agents. 35 Gd(III)<br />

complexes were attached to G6-PAMAM dendrimers along with avidin to afford a targeted delivery<br />

vehicle for NCT. 36 157 Gd provides two advantages. First, it has a large neutron capture cross section<br />

that can be exploited in NCT, and second, Gd(III) complexes can be used as MRI contrast agents. In<br />

fact, dendrimers bearing Gd(III) complexes on their periphery have undergone clinical trials as<br />

MRI contrast agents. 37,38 Therefore, the Gd-dendrimer conjugate described above serves a dual<br />

purpose, therapeutic as well as imaging. The Gd-dendrimer complex was also used as a contrast<br />

agent for dynamic MRI to evaluate the perfusion of an antibody viz. herceptin into tumor tissue<br />

because it has similar size (w10 nm). 39 Similarly, folic acid-conjugated 40 and epidermal growth<br />

factor (EGF)-conjugated 41 boronated PAMAM dendrimers have been prepared as targeting agents<br />

in boron NCT. Slow clearance of macromolecular MRI agents is an issue that limits their scope,<br />

particularly in the investigation of tumor angiogenesis. To address this problem, the avidin chase<br />

system was employed where biotinylated dendrimeric Gd-complexes were rapidly cleared<br />

(w2 min) from blood pool to the liver on injection of avidin. 42 Dendrimeric Gd(III) chelates<br />

with high relaxivity and fast water exchange have been developed to improve upon the currently<br />

available gadolinium complexes that have slow water exchange rates. 43 In rare examples of interior<br />

functionalized dendrimers, boron clusters have been incorporated in the intermediate layers of<br />

water-soluble dendrimers for application in NCT. 44,45<br />

Porphyrins and phthalocyanines are two important macrocycles for use as photosensitizers in<br />

PDT because of large p-conjugation domains and high singlet oxygen quantum yields. 21 However,<br />

large p-conjugation and hydrophobicity in these molecules also lead to a tendency to aggregate that<br />

reduces their solubility in water and precludes use in biological systems. The aggregation also<br />

causes self-quenching and decreases the efficiency of singlet oxygen formation. Incorporation of a<br />

porphyrin or phthalocyanine chromophore at the core of a benzyl ether dendrimer has been shown<br />

to provide the site isolation and to necessarily prevent the aggregation and self-quenching. Hydrophilic<br />

groups such as oligoethylene glycol were attached to the periphery of these dendrimers,<br />

FA<br />

Oligonucleotide<br />

FIGURE 26.4 PAMAM clusters with multiple functionalities prepared via DNA hybridization.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

FA


514<br />

making them useful in biological applications. 46 Oligoethylene glycols are important in biological<br />

applications because they provide solubility in water, reduce hepatic uptake, and enhance biodistribution.<br />

It is to be noted that benzyl ether dendrimers are another class of dendrimers, apart from<br />

PAMAM dendrimers, that can be readily synthesized and have been widely studied for many<br />

applications. Attachment of various functionalities and drugs on periphery and in the interior<br />

can also be carried out with benzyl ether dendrimers using recently developed synthetic<br />

strategies. 47–50<br />

One of the current goals in PDT is to design porphyrin sensitizers that can absorb in the near<br />

infra-red (NIR) region. This is highly desirable because skin is more transparent to light of this<br />

wavelength. However, porphyrins have low two-photon absorption (TPA) cross-sections that limits<br />

their use in singlet oxygen generation. Direct chemical modification of the macrocycles has been<br />

attempted to enhance the TPA cross-section and water solubility. 21 An alternate strategy using<br />

dendrimers was recently developed where donor chromophores with a high TPA cross-section were<br />

covalently attached to the periphery of a benzyl ether dendrimer, containing a porphyrin moiety at<br />

the core. 51 It was observed that emission from porphyrin moiety in the dendritic derivative was 17<br />

times higher than from a porphyrin without the dendritic donor substituents. These dendrimers were<br />

then decorated with water-soluble triethyleneglycol moieties to further expand the scope of their<br />

application. 52 In another study, benzyl ether dendrimers with porphyrin unit at the core were made<br />

water-soluble <strong>by</strong> attachment of quaternary ammonium or carboxylic groups (Figure 26.5), and their<br />

efficiency as photosensitizers was evaluated. 53 To further enhance water-solubility and cellular<br />

uptake of the dendrimer porphyrins with charged periphery, poly-ion complex (PIC) micelles<br />

composed of polyethylene glycol–poly(L-lysine) block copolymers were developed that could<br />

entrap the negatively charged dendrimers through electrostatic association (Figure 26.6). 54 These<br />

polymeric micelle-based nanocarriers showed enhanced cellular uptake, higher photocytotoxicity,<br />

and similar singlet oxygen generation compared to free dendrimer porphyrins. The higher cellular<br />

uptake was due to shielding of negatively charged dendrimer periphery inside a charge-neutral<br />

micelle. The PIC micelles carrying anionic dendrimers were further tested in vivo for treatment of<br />

age-related macular degeneration (AMD) caused <strong>by</strong> choroidal neovascularization (CNV). 55 Experimental<br />

CNV lesions were created in rats <strong>by</strong> laser photocoagulation, and PIC micelles were injected<br />

after seven days. The dendrimer-loaded micelles were found to accumulate preferentially in CNV<br />

lesions and were retained even after 24 h, whereas free dendrimers were cleared within 24 h. When<br />

cationic dendrimer porphyrins were encapsulated in the polyethylene glycol–poly(aspartic acid)<br />

block copolymers, the cellular uptake was lower, and the photodynamic efficacy was higher<br />

compared to free dendrimers. 56 In all these PIC micelles, the dark systemic toxicity (cytotoxicity<br />

of dendrimers under dark conditions) was found to be reduced probably as a result of a biocompatible<br />

poly(ethylene glycol) (PEG) shell.<br />

In a different approach to PDT, 5-aminolevulinic acid (ALA), a natural precursor of protoporphyrin<br />

IX, was attached to the periphery of first and second generation dendrimers via ester<br />

linkages to obtain dendrimer–ALA prodrugs (Figure 26.7). 57 ALA can be intracellularly released<br />

<strong>by</strong> non-specific esterases leading to higher intracellular ALA levels and higher protoporphyrin IX<br />

concentration. Also, the dendrimer–ALA conjugate is more lipophilic than only ALA and can<br />

easily cross the cell membranes as was evident <strong>by</strong> higher uptake of the second-generation<br />

dendrimer studied using tumorigenic cell line PAM 212 in vitro. These dendrimers also showed<br />

negligible dark systemic toxicity.<br />

26.3 ENCAPSULATION OF THE DRUG IN A DENDRIMER<br />

Nanotechnology for Cancer Therapy<br />

Most of the anti-cancer drugs are hydrophobic, and their entrapment in a hydrophilic dendrimer is a<br />

useful strategy to enhance their solubility for drug delivery. This kind of drug carriers composed of<br />

dendrimeric micelles have been a subject of a recent review 12 and is discussed here only briefly.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendritic Nanostructures for Cancer Therapy 515<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

X<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

N<br />

X<br />

O<br />

X<br />

O<br />

N<br />

N<br />

O O<br />

Physical encapsulation prevents the drug from being prematurely degraded. The periphery of the<br />

dendrimer is rendered hydrophilic, either <strong>by</strong> attachment of oligoethylene glycol chains or charged<br />

functional groups such as carboxylates, and the hydrophobic interior is suitable for sequestration of<br />

hydrophobic molecules. 58–61 The drug loading capacity in a dendrimer <strong>by</strong> encapsulation depends<br />

upon its generation and size; however, higher payloads can be obtained only <strong>by</strong> attaching the drug<br />

molecules to the peripheral groups. Several examples of modified PAMAM dendrimers used for<br />

physical encapsulation of drugs have been reported. 62,63<br />

Melamine-based dendrimers modified on the surface with cationic, anionic, and oligoethylene<br />

glycol groups have been used to encapsulate methotrexate and 6-mercaptopurine, and their hepatotoxicity<br />

has been studied in vivo. 64,65 A synthetic strategy has been developed for these<br />

dendrimers as well to incorporate multiple functionalities on the periphery. A hydrophobic anticancer<br />

drug, 10-hydroxycamptothecin, was encapsulated in hydrophilic poly(glycerol succinic<br />

acid) (PGLSA) dendrimers. 66 Cytotoxicity assays of these drug-loaded dendrimers on human<br />

breast cancer cells showed that the drug retained its activity upon encapsulation.<br />

N<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X X<br />

FIGURE 26.5 A water-soluble benzyl ether dendrimer with porphyrin moiety at the core for PDT.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Zn<br />

X<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

X<br />

O<br />

O<br />

O<br />

O<br />

X<br />

O<br />

O<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X<br />

X = COOH


516<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

x<br />

O<br />

O<br />

x<br />

O<br />

x<br />

O<br />

O O<br />

x<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

x<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

x<br />

O O<br />

H<br />

N Zn N<br />

H<br />

x x<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O O O<br />

O<br />

x x x x<br />

O<br />

O O<br />

O<br />

O<br />

O<br />

Negatively charged dendrimer<br />

with a porphyrin core<br />

O<br />

O<br />

O<br />

O<br />

O<br />

x<br />

O<br />

O<br />

x<br />

O<br />

x<br />

O<br />

O<br />

O<br />

O<br />

x<br />

O<br />

O<br />

H<br />

N<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

x<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

NH<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

PEG-b-PLL<br />

HN<br />

O<br />

N<br />

H<br />

O<br />

O<br />

HN<br />

O<br />

O<br />

O<br />

H<br />

N<br />

NH<br />

O NH<br />

FIGURE 26.7 Boc-protected ALA–dendrimer conjugate for PDT.<br />

O<br />

O<br />

HN<br />

O<br />

O<br />

HN<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HN<br />

O<br />

O<br />

O<br />

O<br />

Poly-ion complex (PIC) micelle<br />

FIGURE 26.6 A cartoon showing poly-ion complex (PIC) micelles encapsulating dendrimer–porphyrins.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O HN<br />

O<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

NH<br />

O<br />

O<br />

O


Dendritic Nanostructures for Cancer Therapy 517<br />

26.4 POLYMERIC DRUG CARRIERS BASED ON DENDRIMERS<br />

Dendrimers have been explored as building blocks in the construction of macromolecular scaffolds<br />

for drug delivery to circumvent the following problems: rapid release of drugs physically encapsulated<br />

in a dendrimer, and shorter circulation time of dendrimers themselves. The dendrimer<br />

component in linear–dendritic hybrid polymers provides the large number of functionalities<br />

required for high payload, whereas the linear component affords high molecular weight unattainable<br />

in dendrimers. Such a combination of desirable properties in these hybrid polymers is expected<br />

to provide higher plasma half-life (circulation time) and enhanced cellular uptake because of EPR<br />

effect and high potency. With this premise, dendritic aliphatic polyesters to which doxorubicin<br />

(DOX) was attached via an acid-labile hydrazone linker were connected to a three-arm polyethylene<br />

glycol star polymer to obtain high molecular weight hybrid polymers (Figure 26.8). 67 This<br />

polymer–drug conjugate showed less cytotoxicity (human breast cancer cell lines MDA-MD-231<br />

and MDA-MD-435 using sulforhodamine B assay) than the free drug and longer circulation time<br />

(tested using 125 I radiolabeling) than only dendrimers. These dendritic aliphatic polyesters were<br />

also used to prepare linear–dendritic bow-tie hybrids <strong>by</strong> attaching PEG chains to dendrimer periphery<br />

(illustrated <strong>by</strong> a schematic in Figure 26.9) and evaluated for biodistribution. 68 Bow-tie polymers<br />

with molecular weights up to 1,60,000 g/mol were synthesized, and those with higher molecular<br />

weight (O40,000 g/mol) and more number of linear arms were studied. When tested on mice<br />

bearing subcutaneous B16F10 tumors, these polymers were found to accumulate mainly in<br />

R<br />

OH<br />

R<br />

N<br />

R<br />

OH<br />

N<br />

OH<br />

NH<br />

N<br />

O<br />

NH<br />

O<br />

O<br />

R<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

N<br />

OH<br />

O<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

HO<br />

O<br />

n<br />

O<br />

O<br />

O O O<br />

O<br />

n<br />

O O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

N O<br />

R<br />

N<br />

O<br />

O<br />

O O<br />

O<br />

NH NH<br />

HO<br />

R HO<br />

N<br />

N<br />

R<br />

R HO<br />

NH<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

n<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

O<br />

NH<br />

O<br />

O<br />

O<br />

O<br />

NH<br />

R<br />

NH N<br />

NH<br />

N R<br />

HO<br />

N<br />

N<br />

H2N O<br />

O<br />

OH<br />

CH3 OH O OCH3 OH O<br />

FIGURE 26.8 A linear–dendritic hybrid star polymer with doxorubicin molecules attached to the periphery<br />

via acid-labile hydrazone linkages.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

R =<br />

DOX<br />

HO<br />

HO<br />

OH<br />

R<br />

R


518<br />

tumors with little concentration in the liver, kidney, and lungs. Star-like polymers containing DOX<br />

have also been synthesized with PAMAM dendrimer as the core. For example, starlike poly-N-<br />

(2-hydroxypropyl) methacrylamide (HPMA) was synthesized <strong>by</strong> attachment of the copolymers to<br />

PAMAM dendrimers of generation 2, 3, and 4. 69 DOX was attached to the methacrylamide backbone<br />

via a peptide linker (GlyPheLeuGly) because this linker has been shown to get cleaved <strong>by</strong> an<br />

enzyme Cathepsin B that is over-expressed in malignant cells. 70 The effect of molecular architecture<br />

on release of DOX in vitro was clearly evident <strong>by</strong> the slower release of DOX from the star-like<br />

polymer compared to that from the linear polymer.<br />

26.5 SUMMARY<br />

Dendrimers have clearly emerged as important tools in drug delivery systems. Although tedious<br />

synthesis of dendrimers hampers their widespread use in biological applications, studies on<br />

commercially available dendrimers have clearly demonstrated their potential in drug delivery.<br />

Toward realizing their full potential, however, a combination of useful structural features of<br />

dendrimers with those of well-established biocompatible linear polymers is necessary. This will<br />

provide the key to address important issues involving nanoscale medical devices in cancer therapy.<br />

REFERENCES<br />

Drug moiety<br />

Dendrimer<br />

Nanotechnology for Cancer Therapy<br />

Linear polymer<br />

FIGURE 26.9 Structure of a dendritic–linear bow-tie hybrid polymer is illustrated with a cartoon.<br />

1. Hedrick, J. L. et al., Application of complex macromolecular architectures for advanced microelectronic<br />

materials, Chem. Eur. J., 8, 3308, 2002.<br />

2. Son, S. J. et al., Magnetic nanotubes for magnetic-field-assisted bioseparation, biointeraction, and<br />

drug delivery, J. Am. Chem. Soc., 127, 7316, 2005.<br />

3. Weissleder, R. et al., Cell-specific targeting of nanoparticles <strong>by</strong> multivalent attachment of small<br />

molecules, Nat. Biotechnol., 23, 1418, 2005.<br />

4. Katz, E. and Willner, I., Nanobiotechnology: Integrated nanoparticle–biomolecule hybrid systems:<br />

Synthesis, properties, and applications, Angew. Chem. Int. Ed., 43, 6042, 2004.<br />

5. Duncan, R., The dawning era of polymer therapeutics, Nat. Rev. Drug Discov., 2, 347, 2003.<br />

6. Gillies, E. R. and Frechet, J. M. J., Dendrimers and dendritic polymers in drug delivery, Drug Discov.<br />

Today, 10, 35, 2005.<br />

7. Kim, Y. and Zimmerman, S. C., Applications of dendrimers in bio-organic chemistry, Curr. Opin.<br />

Chem. Biol., 2, 733, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendritic Nanostructures for Cancer Therapy 519<br />

8. Boas, U. and Heegaard, P. M. H., Dendrimers in drug research, Chem. Soc. Rev., 33, 43, 2004.<br />

9. Tully, D. C. and Frechet, J. M. J., Dendrimers at surfaces and interfaces: Chemistry and applications,<br />

Chem. Commun., 14, 1229, 2001.<br />

10. Patri, A. K., Majoros, I. J., and Baker, J. R., Dendritic polymer macromolecular carriers for drug<br />

delivery, Curr. Opin. Chem. Biol., 6, 466, 2002.<br />

11. Haag, R. and Vogtle, F., Highly branched macromolecules at the interface of chemistry, biology,<br />

physics, and medicine, Angew. Chem. Int. Ed., 43, 272, 2004.<br />

12. Ambade, A. V., Savariar, E. N., and Thayumanavan, S., Dendrimeric micelles for controlled drug<br />

release and targeted delivery, Mol. Pharm., 2, 264, 2005.<br />

13. Bosman, A. W., Janssen, H. M., and Meijer, E. W., About dendrimers: Structure, physical properties,<br />

and applications, Chem. Rev., 99, 1665, 1999.<br />

14. Basu, S., Sandanaraj, B. S., and Thayumanavan, S., Molecular recognition in dendrimers, In Encyclopedia<br />

of Polymer Science and Technology, 4th ed., Mark, H. F., Ed., Wiley, New York, 2004.<br />

15. Grayson, S. M. and Frechet, J. M. J., Convergent dendrons and dendrimers: From synthesis to<br />

applications, Chem. Rev., 101, 3819, 2001.<br />

16. Newkome, G. R., Moorefield, C. N., and Vogtle, F., Eds., Dendrimers and Dendrons: Concepts,<br />

Syntheses, Applications, 2nd ed., Wiley-VCH, Weinheim, 2001.<br />

17. Abu-Rmaileh, R., Attwood, D., and D’Emanuele, A., Dendrimers in cancer therapy, Drug Deliv. Syst.<br />

Sci., 3, 65, 2003.<br />

18. Sampathkumar, S. -G. and Yarema, K. J., Targeting cancer cells with dendrimers, Chem. Biol., 12, 5,<br />

2005.<br />

19. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and antitumor agent smancs,<br />

Cancer Res., 46, 6387, 1986.<br />

20. Henderson, B. W. and Dougherty, T. J., Eds., Photodynamic Therapy: Basic Principles and Clinical<br />

Applications, Marcel Dekker, New York, 1992.<br />

21. Lunardi, C. N. and Tedesco, A. C., Synergic photosensitizers: A new trend in photodynamic therapy,<br />

Curr. Org. Chem., 9, 813, 2005.<br />

22. Barth, R. F. et al., Boron neutron capture therapy of cancer: Current status and future prospects, Clin.<br />

Cancer Res., 11, 3987, 2005.<br />

23. Matthews, S. E., Pouton, C. W., and Threadgill, M. D., Macromolecular systems for chemotherapy<br />

and magnetic resonance imaging, Adv. Drug Deliv. Rev., 18, 219, 1996.<br />

24. De Groot, F. M. H. et al., Cascade-release dendrimers liberate all end groups upon a single triggering<br />

event in the dendritic core, Angew. Chem. Int. Ed., 42, 4490, 2003.<br />

25. Esfand, R. and Tomalia, D. A., Poly(amidoamine) (PAMAM) dendrimers: From biomimicry to drug<br />

delivery and biomedical applications, Drug Discov. Today, 6, 427, 2001.<br />

26. Malik, N. et al., Dendrimers: Relationship between structure and biocompatibility in vitro, and<br />

preliminary studies on the biodistribution of 125 I labeled polyamidoamine dendrimers in vivo,<br />

J. Control. Release, 65, 133, 2000.<br />

27. Shi, X., Majoros, I. J., and Baker, J. R., Capillary electrophoresis of poly(amidoamine) dendrimers:<br />

From simple derivatives to complex multifunctional medical nanodevices, Mol. Pharm., 2, 278, 2005.<br />

28. Khandare, J. et al., Synthesis, cellular transport and activity of polyamidoamine dendrimer-methylprednisolone<br />

conjugates, Bioconjug. Chem., 16, 330, 2005.<br />

29. Patri, A. K. et al., Synthesis and in vitro testing of J591 antibody–dendrimer conjugate for targeted<br />

prostate cancer therapy, Bioconjug. Chem., 15, 1174, 2004.<br />

30. Majoros, I. J. et al., Poly(amidoamine) dendrimer-based multifunctional engineered nanodevice for<br />

cancer therapy, J. Med. Chem., 48, 5892, 2005.<br />

31. Choi, Y. et al., Synthesis and functional evaluation of DNA-assembled polyamidoamine dendrimer<br />

clusters for cancer cell-specific targeting, Chem. Biol., 12, 35, 2005.<br />

32. Bezouska, K., Design, functional evaluation and biomedical applications of carbohydrate dendrimers<br />

(glycodendrimers), Rev. Mol. Biotechnol., 90, 269, 2002.<br />

33. Woller, E. K. et al., Altering the strength of lectin binding interactions and controlling the amount of<br />

lectin clustering using mannose/hydroxyl-functionalized dendrimers, J. Am. Chem. Soc., 125, 8820,<br />

2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


520<br />

Nanotechnology for Cancer Therapy<br />

34. Page, D. and Roy, R., Synthesis and biological properties of mannosylated starburst poly(amidoamine)<br />

dendrimers, Bioconjug. Chem., 8, 714, 1997.<br />

35. Venditto, V. J. et al., PAMAM dendrimer based macromolecules as improved contrast agents, Mol.<br />

Pharm., 2, 302, 2005.<br />

36. Kobayashi, H. et al., Avidin–dendrimer–(1B4M-Gd)254: A tumor-targeting therapeutic agent for<br />

gadolinium neutron capture therapy of intraperitoneal disseminated tumor which can be monitored<br />

<strong>by</strong> MRI, Bioconjug. Chem., 12, 587, 2001.<br />

37. Wiener, E. and Narayanan, V. V., Magnetic resonance imaging contrast agents: Theory and the role of<br />

dendrimers, Adv. Dendr. Macromol., 5, 129, 2002.<br />

38. Stiriba, S. E., Frey, H., and Haag, R., Dendritic polymers in biomedical applications: From potential to<br />

clinical use in diagnostics and therapy, Angew. Chem. Int. Ed., 41, 1329, 2002.<br />

39. Kobayashi, H. et al., Rapid accumulation and internalization of radiolabeled herceptin in an inflammatory<br />

breast cancer xenograft with vasculogenic mimicry predicted <strong>by</strong> the contrast-enhanced<br />

dynamic MRI with the macromolecular contrast agent G6-(1B4M-Gd)256, Cancer Res., 62, 860, 2002.<br />

40. Shukla, S. et al., Synthesis and biological evaluation of folate receptor-targeted boronated PAMAM<br />

dendrimers as potential agents for neutron capture therapy, Bioconjug. Chem., 14, 158, 2003.<br />

41. Capala, J. et al., Boronated epidermal growth factor as a potential targeting agent for boron neutron<br />

capture therapy of brain tumors, Bioconjug. Chem., 7, 7, 1996.<br />

42. Kobayashi, H. et al., Activated clearance of a biotinylated macromolecular MRI contrast agent from<br />

the blood pool using an avidin chase, Bioconjug. Chem., 14, 1044, 2003.<br />

43. Pierre, V. C., Botta, M., and Raymond, K. N., Dendrimeric gadolinium chelate with fast water<br />

exchange and high relaxivity at high magnetic field strength, J. Am. Chem. Soc., 127, 504, 2005.<br />

44. Newkome, G. R. et al., Chemistry of micelles. 37. Internal chemical transformations in a precursor of a<br />

unimolecular micelle: Boron supercluster via site-specific addition of B10H14 to cascade molecules,<br />

Angew. Chem. Int. Ed. Engl., 33, 666, 1994.<br />

45. Parrott, M. C. et al., Synthesis and properties of carborane-functionalized aliphatic polyester dendrimers,<br />

J. Am. Chem. Soc., 127, 12081, 2005.<br />

46. Brewis, M. et al., Phthalocyanine-centred aryl ether dendrimers with oligo(ethyleneoxy) surface<br />

groups, Tetrahedron Lett., 42, 813, 2001.<br />

47. Freeman, A. W., Chrisstoffels, L. A. J., and Frechet, J. M. J., A Simple method for controlling<br />

dendritic architecture and diversity: A parallel monomer combination approach, J. Org. Chem., 65,<br />

7612, 2000.<br />

48. Sivanandan, K., Sandanaraj, B. S., and Thayumanavan, S., Sequences in dendrons and dendrimers,<br />

J. Org. Chem., 69, 2937, 2004.<br />

49. Vutukuri, D., Sivanandan, K., and Thayumanavan, S., Synthesis of dendrimers with multifunctional<br />

periphery using an ABB 0 monomer, Chem. Commun., 21, 796–797, 2003.<br />

50. Sivanandan, K., Vutukuri, D., and Thayumanavan, S., Functional group diversity in dendrimers, Org.<br />

Lett., 4, 3751, 2002.<br />

51. Dichtel, W. R. et al., Singlet oxygen generation via two-photon excited FRET, J. Am. Chem. Soc., 126,<br />

5380, 2004.<br />

52. Oar, M. A. et al., Photosensitization of singlet oxygen via two-photon excited fluorescence resonance<br />

energy transfer in a water-soluble dendrimer, Chem. Mater., 17, 2267, 2005.<br />

53. Nishiyama, N. et al., Light-harvesting ionic dendrimer porphyrins as new photosensitizers for photodynamic<br />

therapy, Bioconjug. Chem., 14, 58, 2003.<br />

54. Jang, W. -D. et al., Supramolecular nanocarrier of anionic dendrimer porphyrins with cationic block<br />

copolymers modified with polyethylene glycol to enhance intracellular photodynamic efficacy,<br />

Angew. Chem. Int. Ed., 44, 419, 2005.<br />

55. Ideta, R. et al., Nanotechnology-based photodynamic therapy for neovascular disease using a supramolecular<br />

nanocarrier loaded with a dendritic photosensitizer, Nano Lett., 5(12), 2426–2431, 2005.<br />

56. Zhang, G. -D. et al., Polyion complex micelles entrapping cationic dendrimer porphyrin: Effective<br />

photosensitizer for photodynamic therapy of cancer, J. Control Release, 93, 141, 2003.<br />

57. Battah, S. H. et al., Synthesis and biological studies of 5-aminolevulinic acid-containing dendrimers<br />

for photodynamic therapy, Bioconjug. Chem., 12, 980, 2001.<br />

58. Vutukuri, D. R., Basu, S., and Thayumanavan, S., Dendrimers with both polar and apolar nanocontainer<br />

characteristics, J. Am. Chem. Soc., 126, 15636, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendritic Nanostructures for Cancer Therapy 521<br />

59. Gopidas, K. R., Whitesell, J. K., and Fox, M. A., Metal-core-organic shell dendrimers as unimolecular<br />

micelles, J. Am. Chem. Soc., 125, 14168, 2003.<br />

60. Hawker, C. J., Wooley, K. L., and Frechet, J. M. J., Unimolecular micelles and globular amphiphiles:<br />

Dendritic macromolecules as novel recyclable solubilization agents, J. Chem. Soc. Perkin Trans., 1,<br />

1287, 1993.<br />

61. Aathimanikandan, S. V., Savariar, E. N., and Thayumanavan, S., Temperature-sensitive dendritic<br />

micelles, J. Am. Chem. Soc., 127, 14922, 2005.<br />

62. Beezer, A. E. et al., Dendrimers as potential drug carriers; encapsulation of acidic hydrophobes within<br />

water soluble PAMAM derivatives, Tetrahedron, 59, 3873, 2003.<br />

63. Kojima, C. et al., Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and<br />

their ability to encapsulate anticancer drugs, Bioconjug. Chem., 11, 910, 2000.<br />

64. Neerman, M. F. et al., Reduction of drug toxicity using dendrimers based on melamine, Mol. Pharm.,<br />

1, 390, 2004.<br />

65. Zhang, W. et al., Orthogonal, convergent syntheses of dendrimers based on melamine with one or two<br />

unique surface sites for manipulation, J. Am. Chem. Soc., 123, 8914, 2001.<br />

66. Morgan, M. T. et al., Dendritic molecular capsules for hydrophobic compounds, J. Am. Chem. Soc.,<br />

125, 15485, 2003.<br />

67. Padilla De Jesus, O. L. et al., Polyester dendritic systems for drug delivery applications: In vitro and<br />

in vivo evaluation, Bioconjug. Chem., 13, 453, 2002.<br />

68. Gillies, E. R. et al., Biological evaluation of polyester dendrimer: Poly(ethylene oxide) “bow-tie”<br />

hybrids with tunable molecular weight and architecture, Mol. Pharm., 2, 129, 2005.<br />

69. Wang, D. et al., Synthesis of starlike N-(2-hydroxypropyl) methacrylamide copolymers: Potential<br />

drug carriers, Biomacromolecules, 1, 313, 2000.<br />

70. Fuchs, S. et al., Fluorescent dendrimers with a peptide cathepsin B cleavage site for drug delivery<br />

applications, Chem. Commun., 14, 1830–1832, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


27<br />

CONTENTS<br />

PEGylated Dendritic<br />

Nanoparticulate Carriers of<br />

Anti-Cancer Drugs<br />

D. Bhadra, S. Bhadra, and N. K. Jain<br />

27.1 Introduction ....................................................................................................................... 523<br />

27.2 Problems in Cancer Chemotherapy .................................................................................. 524<br />

27.3 Nanoparticles in Cancer Chemotherapy ........................................................................... 525<br />

27.3.1 Challenges of Nanotechnology ........................................................................... 527<br />

27.4 Poly(Ethylene Glycol) Conjugation .................................................................................. 527<br />

27.4.1 Derivatization and Activation of Poly(Ethylene Glycol) ................................... 528<br />

27.4.2 Properties of Poly(Ethylene Glycol) Conjugates................................................ 528<br />

27.5 Dendrimer as Drug Delivery System: Selection, Applicability, and Rationality ............ 529<br />

27.5.1 Differences Between Hyperbranched Structures and Dendrimers ..................... 532<br />

27.5.2 Classification of Dendrimers............................................................................... 532<br />

27.5.3 Features of Dendrimers as Nanodrugs................................................................ 533<br />

27.6 Applications of PEGylated Dendrimers............................................................................ 534<br />

27.6.1 Enhanced Biocompatibility for Drug Delivery .................................................. 535<br />

27.6.2 Micellar Means of Enhancing Solubility............................................................ 537<br />

27.6.3 Enhancing Pharmacokinetic Properties .............................................................. 538<br />

27.6.4 Dendrimer Use as Transfection Reagents .......................................................... 539<br />

27.6.5 Stimuli Responsive Carriers................................................................................ 540<br />

27.6.6 Alteration of Toxicity.......................................................................................... 541<br />

27.6.7 Effective Photodynamic Therapy........................................................................ 543<br />

27.6.8 Dendritic Gels ..................................................................................................... 543<br />

27.6.9 Ligand-Based Drug Delivery .............................................................................. 543<br />

27.6.10 Increasingly Effective Boron Neutron Capture Therapy.................................... 544<br />

27.6.11 Improved MRI Contrast Agents.......................................................................... 544<br />

27.7 Conclusions ....................................................................................................................... 545<br />

References .................................................................................................................................... 545<br />

27.1 INTRODUCTION<br />

Cancer is a leading cause of death. Irrespective of etiology, cancer is basically a disease of cells<br />

characterized <strong>by</strong> loss of normal cellular growth, maturation, and multiplication that lead to disturbance<br />

of homeostasis. The main features of cancer are (Barar 2000) excessive cell growth,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

523


524<br />

invasiveness, undifferentiated cells or tissues, the ability to metastasize or spread to newer sites and<br />

establish new growths, a type of acquired heredity where the progeny of cancer cells also retain<br />

cancerous properties, and a shift of cellular metabolism, leading to increased production of macromolecules<br />

from nucleosides and amino acids that causes an increased metabolism of carbohydrates<br />

for cellular energy, ultimately resulting in the host’s illness. This illness is attributed to pressure<br />

effects as a result of local tumor growth, destruction of the organs involved <strong>by</strong> the primary growth<br />

or its metastases, and systemic effects as a result of the new growths.<br />

A single cancerous cell surrounded <strong>by</strong> healthy tissue will replicate at a rate higher than the other<br />

cells, placing a strain on the nutrient supply and elimination of metabolic waste products. Tumor cells<br />

will displace healthy cells until the tumor reaches a diffusion-limited maximal size. Although tumor<br />

cells will typically not initiate apoptosis in a low nutrient environment, they do require the normal<br />

building blocks of cell function like oxygen, glucose, and amino acids. The vasculature that was<br />

designed to supply the now-extinct healthy tissue could not place as high a demand for nutrients<br />

because of its slower growth rate. Tumor cells will continue dividing because they do so without<br />

regard to nutrient supply, but many tumor cells will perish because the amount of nutrients is<br />

insufficient. The tumor cells at the outer edge of a mass have the best access to nutrients, and cells<br />

on the inside die creating a necrotic core within tumors that rely on diffusion to deliver nutrients and<br />

eliminate waste products. In essence, a steady state tumor size forms as the rate of proliferation is<br />

equal to the rate of cell death until a better connection with the circulatory system is created. This<br />

diffusion-limited maximal size of most tumors is around 2 mm 3 (Grossfeld, Carrol, and Lindeman<br />

2002). To grow beyond this size, the tumor must recruit the formation of blood vessels to provide the<br />

necessary nutrients to fuel its continued expansion. It is thought that there could be numerous tumors<br />

at this diffusion-limited maximal size throughout the body. Until the tumor can gain that access to the<br />

circulation, it will remain at this size, and the process can take years.<br />

27.2 PROBLEMS IN CANCER CHEMOTHERAPY<br />

Nanotechnology for Cancer Therapy<br />

In the past, chemotherapy was considered as a last resort after more successful treatments like<br />

surgery and radiotherapy had failed. The main problem of cancer chemotherapy is the lack of<br />

highly selective drugs, and the rapidly dividing normal cells of the bone marrow, gut, lymphoid<br />

tissue, spermatogenic cells, fetus, and hair follicles are also killed. Most of the antineoplastic drugs<br />

act on the processes such as DNA synthesis, transcription, or the mitotic phase, and they are labeled<br />

as cell cycle phase-specific drugs (also known as phase-dependent drugs). The phase-specific drugs<br />

do not act on G 0 phase. In contrast, there are certain drugs that kill the cells during all or most<br />

phases of the cycle, and they are labeled as cell cycle phase-nonspecific drugs (also known as<br />

phase-independent drugs). The phase-specific drugs have proven to be effective in hematological<br />

malignancies and tumors with high rate of proliferation or high growth fraction (Barar 2000).<br />

The various obstacles for drug delivery to tumors include differences in cellular morphology,<br />

tissue immunogenecity, uncontrolled rate of growth, capacity of metastasize, and poor response to<br />

chemotherapeutic agents. Furthermore, there are variations in blood flow and vessel permeability<br />

within different regions of some tissues (Reynolds 1996). Various drug delivery systems are<br />

employed for delivery of chemotherapeutic agents to neoplastic cells. This helps in targeting<br />

drugs directly into cells and preventing drug interaction with normal tissues and alleviating side<br />

effects to normal cells. Because of a lack of highly selective drugs for tumor cells and tissues and<br />

decreased penetration of drug into neoplastic cells, a number of novel drug delivery systems,<br />

prodrugs, and other chemically modified forms of drugs having altered tissue distributions were<br />

brought in use and recently reviewed for targeting to cells located in various parts of body (Brigger,<br />

Dubernet, and Couvreur 2002).<br />

There are various means to effectively fight the tumors such as achieving targeting <strong>by</strong><br />

avoiding reticuloendothelial system (RES); targeted delivery through enhanced permeability<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 525<br />

and retention (EPR) effects; tumor-specific targeting; targeting through angiogenesis; targeting<br />

tumor vasculaturel; etc. (Brannon-Peppas and Blanchette 2004). Particles with longer circulation<br />

times and greater ability to target to the site of interest should be 100 nm or less in diameter and<br />

have a hydrophilic surface in order to reduce clearance <strong>by</strong> macrophages (Storm et al. 1995).<br />

Coatings of hydrophilic polymers can create a cloud of chains at the particle surface that will<br />

repel plasma proteins, and work in this area began <strong>by</strong> adsorbing surfactants to the<br />

nanoparticles surface.<br />

Other routes include forming the particles from branched or block copolymers with hydrophilic<br />

and hydrophobic domains. One potential advantage in treatment of advance stage cancerous tissue<br />

is the inherent leaky vasculature that allows for greater accumulation of colloidal systems. The<br />

check the defective vascular architecture, created as a result of the rapid vascularization necessary<br />

to serve fast-growing cancers, coupled with poor lymphatic drainage allows an enhanced permeation<br />

and retention effect (EPR effect) (Sledge and Miller 2003). The ability to target<br />

treatment to very specific cancer cells also uses a cancer’s own structure in that many cancers<br />

over express particular antigens, even on their surface. This makes them ideal targets for drug<br />

delivery as long as the targets for a particular cancer cell type can be identified with confidence and<br />

are not expressed in significant quantities anywhere else in the body. Tumor-activated prodrug<br />

therapy uses the approach that a drug conjugated to a tumor-specific molecule will remain inactive<br />

until it reaches the tumor (Chari 1998). These systems would ideally be dependent on interactions<br />

with cells found specifically on the surface of cancerous cells and not the surface of healthy cells.<br />

Most linkers are usually peptidase cleavable or acid labile but may not be stable enough in vivo to<br />

give desirable clinical outcomes.<br />

Limitations also exist because of the lower potency of some drugs after being linked to<br />

targeting moieties when the targeting portion is not cleaved correctly or at all. For example,<br />

adriamycin-conjugated poly(ethylene glycol) linker with enzymatically cleavable peptide<br />

sequences (alanyl-valine, alanyl-proline, and glycyl-proline) or using monoclonal antibodies has<br />

shown a greater selectivity to cleavage at tumor cells (Suzawa et al. 2002). A number of targeted<br />

cancer treatments using antibodies for specific cancer types have been approved <strong>by</strong> the U.S. Food<br />

and Drug Administration like rituximab, trastuzumab, gemtuzumabozogamicin, alemtuzumab,<br />

ibritumomab tiuxetan, gefitinib, etc. (Abou-Jawde et al. 2003).<br />

27.3 NANOPARTICLES IN CANCER CHEMOTHERAPY<br />

Nanotechnology applied to cancer treatment may offer several promising advantages over conventional<br />

drugs. Nanoscale devices are two orders of magnitude smaller than tumor cells, making it<br />

possible for them to directly interact with intracellular organelles and proteins. Because of their<br />

molecule-like size, nanoscale tools may be capable of early disease detection using minimal amounts<br />

of tissue, even down to a single malignant cell (NCI 2005). These tools may not only prevent disease<br />

<strong>by</strong> monitoring genetic damage, but also treat cells in vivo while minimizing interference with healthy<br />

tissue. By combining different kinds of nanoscale tools on a single device, it may be possible to run<br />

multiple diagnostic tests simultaneously (www.math.uci.edu/~cristini/publications/nanochap.pdf).<br />

In particular, it is hoped that cancer drug therapy involving nanotechnology will be more effective in<br />

targeting malignant cells and sparing healthy tissue. In this regard, the role of nanoparticles loaded<br />

with chemotherapeutic drugs has been receiving much attention. Research and development in this<br />

area is expected to dramatically increase in importance in the coming years. In general, nanoscale<br />

drug delivery systems (Figure 27.1) for chemotherapy can be divided into two categories: polymerand<br />

lipid-based (Langer 2000; Sahoo and Labhasetwar, 2003).<br />

Polymers are usually larger than lipid molecules, and they form a solid phase such as polymeric<br />

nanoparticles, films, pellets, dendrimers; whereas, lipids form a liquid (or liquid crystalline phase)<br />

such as liposomes, cubersomes, micelles, and other emulsions (Feng and Chien 2003). Whereas<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


526<br />

ligand<br />

Drug<br />

Drug<br />

Nanosphere Nanocapsule<br />

PEG<br />

ligand<br />

ligand<br />

Liposome<br />

ligand<br />

ZnS<br />

CdSe<br />

ligand<br />

Antibody<br />

ligand<br />

ligand<br />

ligand<br />

Ligand conjugated<br />

ZnS-capped CdSe quantum dots<br />

Hydrophobic<br />

core<br />

Polymeric micelles<br />

polymer-based systems are considered biologically more stable than lipid-based systems, the latter<br />

are generally more biocompatible. Polymer-based systems might possess good drug targeting<br />

ability because their uptake may be different for cells in different tissues (Maruyama 2000). In<br />

fact, Feng and Chien (2003) have suggested that a combination of polymer- and lipid-based systems<br />

could integrate their advantages while avoiding their respective disadvantages. An example of such<br />

a nanoparticle would be a liposomes-in-microspheres (LIM) system where drugs are first loaded<br />

into liposomes and then encapsulated into polymeric microspheres. This way, both hydrophobic<br />

and hydrophilic drugs can be delivered in one nanoparticle. The bioactivity of peptides and proteins<br />

would be preserved in the liposomes whose stability is protected <strong>by</strong> the polymeric matrix (Feng and<br />

Chien 2003).<br />

Chemotherapy using nanoparticles has been studied in clinical trials for several years, and<br />

numerous studies have been published in this regard (Kreuter 1994). Two liposomally<br />

delivered drugs are currently on the market: daunorubicin and doxorubicin (Massing and<br />

Fuxius 2000). These encapsulated drugs can be formulated to maximize their half-life in<br />

the circulation. For example, a stealth version of liposomal doxorubicin coated with polyethylene<br />

glycol to reduce its uptake <strong>by</strong> the RES can extend its half-life in blood for up to 50–60 h<br />

(CCO Formulary 2003).<br />

Core<br />

Nanotechnology for Cancer Therapy<br />

Internal cavity<br />

Branchingunits<br />

Fourth generation dendrimer<br />

Fe3O 4<br />

Ligend<br />

Ferrofluid conjugated to ligand<br />

FIGURE 27.1 Representation of different nanoparticulate carriers used for drug delivery. (Adapted from Sahoo,<br />

S. K. and Labhasetwar, V., Drug Discov. Today, 8, 2003, www.drugdiscoverytoday.com. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 527<br />

27.3.1 CHALLENGES OF NANOTECHNOLOGY<br />

The difficulties faced <strong>by</strong> nanotechnology in the service of clinical medicine are numerous. These<br />

difficulties should be kept in mind while considering chemotherapeutic treatment involving nanotechnology<br />

and the potential role of biocomputation. First, there are basic physical issues that<br />

matter on a small scale. Because matter behaves differently on the nanolevel than it does at<br />

micro- and macro levels, most of the science at the nanoscale has been devoted to basic research,<br />

designed to expand understanding of how matter behaves on this length scale (www.math.uci.edu/~<br />

cristini/publications/nanochap.pdf). Because nanomaterials have large surface areas relative to<br />

their volumes, phenomena such as friction are more critical than they are in larger systems. The<br />

small size of nanoparticles may result in significant delay or speed-up in their intended actions.<br />

They may accumulate at unintended sites in the body. They may provoke unexpected immune<br />

system reactions. Cells may adapt to the nanoparticles, modifying the body’s behavior in unforeseen<br />

ways (www.math.uci.edu/~cristini/publications/nanochap.pdf).<br />

The efficacy of nanoparticles may be adversely affected <strong>by</strong> their interaction with the cellular<br />

environment. For instance, the RES may clear nanoscale devices, even stealth versions, too rapidly<br />

for them to be effective because of the tendency of the RES to phagocytose nanoparticles (Kreuter<br />

1994). Nanoparticles can be taken up <strong>by</strong> dendritic cells (Elamanchili et al. 2004) and <strong>by</strong> macrophages<br />

(Cui, Hsu, and Mumper 2003). RES accumulation of nanoparticles could potentially lead to<br />

a compromise of the immune system. On the other hand, larger nanoparticles may accumulate in<br />

larger organs, leading to toxicity. Perhaps the biggest issue of all is that the physically compromised<br />

tumor vasculature may prevent most of the nanodevices from reaching the target cells <strong>by</strong> vascular<br />

transport or diffusion. Alterations in the tumor vasculature may adversely affect the convection of<br />

the nanodevices in the blood stream (Brigger, Dubernet, and Couvreur 2002). Local cell density and<br />

other stromal features may hamper drug or nanodevice diffusion through tumoral tissue.<br />

27.4 POLY(ETHYLENE GLYCOL) CONJUGATION<br />

Poly(ethylene glycol) is a non-toxic water-soluble polymer that resists recognition <strong>by</strong> the immune<br />

system. The term PEG is used for poly(ethylene glycol) that is often referred for polymer chains<br />

with molecular weight (MWs) below 20,000 Da, and poly(ethylene oxide) (PEO) refers to higher<br />

MW polymers (Harris et al. 1984; Peppas et al. 1999). These are homogeneous polymers of Mw/<br />

Mn!1.1 represented <strong>by</strong> the formula HO(CH2CH2O)nH and is called as poly(oxy-1, 2-ethanediyl),<br />

or a-hydro-u-hydroxy-poly(ethylene glycol). It is an additional polymer of ethylene oxide and<br />

water, where n represents the average number of oxy-ethylene groups. These are generally soluble<br />

in water, alcohol, acetone, and chloroform, but miscible with glycols and practically insoluble in<br />

ether. These are also known as Macrogols (Reynolds 1996) that are generally of higher molecular<br />

weights such as Macrogols of 20,000 Da. The pH of 5% solution is about 4–7.5.<br />

The aqueous solutions of PEGs can be sterilized <strong>by</strong> filtration, autoclaving, etc. and are needed<br />

to be stored in airtight containers. The PEGs are relatively less toxic, but toxicity, if any appears to<br />

be greatest with macrogols of lower molecular weights. On topical administration, it may cause<br />

stinging, especially to mucous membrane. Such administration of PEGs may be associated with<br />

hypersensitivity reactions such as urticaria, and local gastrointestinal discomfort, bloating and<br />

nausea, abdominal cramps, vomiting and irritation, may result on its use in bowel cleansing<br />

preparation. Macrogols demonstrate oxidizing ability, leading to incompatibility with activity of<br />

bacitracin or benzyl-penicillin that may get reduced in macrogol bases. Estimated acceptable daily<br />

intake may be upto 10-mg/kg body weight (Reynolds 1996). Its presence in aqueous solution has no<br />

deleterious effect on protein conformation or activity of enzymes. It exhibits rapid clearance from<br />

the body, and it has been approved for a wide range of biomedical applications. PEG may transfer<br />

its properties to other molecules when it is covalently bound to that molecule. This can result in<br />

modification of number of properties of molecules as illustrated later on (Zalipsky 1995).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


528<br />

PEG has polyether backbone that is inert in biological environment as well as in most chemical<br />

reaction conditions as in chemical modifications and/or conjugation reactions. The chemical derivatisation<br />

of end groups of PEG is an essential first step in preparation of bioconjugates. The<br />

various properties are altered because of polyethylene glycol conjugation that is discussed in the<br />

following sections. PEG conjugation was tried and found most applicable in case of many nanoparticulate<br />

systems, using activated PEGs with appropriate spacers.<br />

27.4.1 DERIVATIZATION AND ACTIVATION OF POLY(ETHYLENE GLYCOL)<br />

Poly(ethylene glycol) has two equivalent hydroxyl groups, so this could act as potential crosslinking<br />

agent for any systems to which it is attached. These hydroxyl groups can be attached to<br />

various bio-active species <strong>by</strong> covalent coupling using number of simple PEG analogues. Some<br />

important stable analogues are bromo, amino, aminoethyl, carboxymethyl, succinimidosuccinate,<br />

tosylate, mesylate, aldehyde, octadecylamine, monopalmitate, stearoyloxy derivative of PEG, or<br />

monomethoxy-PEG (MPEG) (Buckmann and Morr 1981; Harris et al. 1984). These were prepared<br />

for reactions with sensitive bio-active systems under mild conditions.<br />

One very important derivative that has been used in number of derivatisation reactions that has<br />

one hydroxyl group blocked is monomethyl-ether of polyethylene-glycol or monomethoxy-polyethylene-glycol<br />

(MPEG), i.e., considered equivalent in terms of reaction for formation of PEG<br />

derivatives. This was generally brought in use for conjugation to bioactive species. It is generally<br />

used when multiple chains of polymers had to be linked to the intended substrates. Because of<br />

structural simplicity and possession of only one-derivatizable end group, the use of MPEG minimizes<br />

cross-linking possibilities and leads to improved homogeneity of conjugate. Therefore, it<br />

usually is a starting material of choice for the covalent modification of proteins, biomaterials,<br />

particulates, lipids, drugs, etc. (Zalipsky 1995).<br />

MeOH CHðOCH 2CH 2Þ nOH/MeOðCH 2CH 2OÞ nH<br />

Commercially available MPEGs are often contaminated with significant amounts (equivalent to<br />

25%) of dihydroxy-terminated polymers and may be purified <strong>by</strong> chromatography and ethyl ether<br />

precipitation. The use of these polymers as starting materials of choice for covalent modification of<br />

proteins, biomaterials, and particles is dependent for not forming cross-linked conjugates (Harris<br />

et al. 1984).<br />

In most cases, commercially available MPEGs of 2000–5000 MWs are used for preparation of<br />

number of reagents. MPEG-linked and based electrophiles were synthesized and used for linking<br />

with number of available attachment sites, e.g., amino acids and other nucleophilic groups. These<br />

are called activated PEGs (Zalipsky 1995). The intermediate linkers are called spacers, linking to<br />

bioactive groups, especially proteins and enzymes. These generally require mild conditions of<br />

reactions and specific properties of reactants for binding.<br />

In order of reaction and based on ease of introduction, these can roughly be arranged as follows:<br />

arylating agents O acylating agents O alkylating agents (Zalipsky 1995). For example, tresylates,<br />

succinimidyl succinate derivatives, succinimidyl ester of carboxymethyl PEG, oxy-carboxylamino<br />

acid derivative of PEG, acetaldehyde, and hydrazide derivatives were mainly used for PEGylation<br />

of proteins and enzymes.<br />

27.4.2 PROPERTIES OF POLY(ETHYLENE GLYCOL) CONJUGATES<br />

Nanotechnology for Cancer Therapy<br />

This group reviewed various properties affected <strong>by</strong> PEGylation of carriers (Bhadra et al. 2002).<br />

PEGylation has pronounced effects on various parameters of drug delivery systems (Katre 1993;<br />

Zalipsky 1995). As far as bio-distribution and pharmacokinetic, PEG conjugation increases blood<br />

circulation half-life, <strong>by</strong> reducing the tissue distribution, RES, liver, spleen, and macrophage uptake<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 529<br />

of the carriers. (Veronese et al. 1989; Papahadjopoulus et al. 1991; Phillips et al. 1999). PEG<br />

coatings alter solubility of the systems (Choi et al. 1998;Kim et al. 1999). It reduces the toxicity<br />

<strong>by</strong> decreasing the release and contact of active constituent, decreasing immunogenicity, celladherence,<br />

thrombogenicity, and protein adsorption. PEGylation can also lead to stabilization of<br />

delivery systems. It can also improve utilization as drug a delivery system <strong>by</strong> increasing drug<br />

loading per system, improving drug targeting as in case of liposomes <strong>by</strong> diffusion controlling<br />

mechanisms, and promoting its use as a micellar delivery system. This process can sustain and<br />

control drug delivery safely and appropriately <strong>by</strong> decreasing burst effects and <strong>by</strong> decreasing drug<br />

leakage and increasing stability of system. It can improve transfection capability (Choi et al. 1998).<br />

This can increase tumor uptake (Papahadjopoulus et al. 1991; Ishida et al. 1999) of drug and<br />

delivery systems.<br />

Manipulation of the upper layers (corona) of nanoparticles or dendrimers confers advantageous<br />

properties to such particles, e.g., increased solubility and biocompatibility (McNeil<br />

2005). Attaching hydrophilic polymers to the surface such as PEG greatly increases the hydration<br />

(i.e., solubility) of the nanoparticles and can protect attached proteins from enzymatic degradation<br />

when used for in vivo applications (Bhadra et al. 2002; Harris and Chess 2003). Nanoparticles<br />

with hydrophilic polymers such as PEG attached to their surface can act as a platform for<br />

lipophilic molecules, and they can overcome the solubility barrier. Insoluble compounds can<br />

be attached, adsorbed, or otherwise encapsulated in the hydrated nanoparticles. The surface<br />

addition of PEG and other hydrophilic polymers also increases the in vivo compatibility of<br />

nanoparticles. When intravascularly injected, uncoated nanoparticles are rapidly cleared from<br />

the bloodstream <strong>by</strong> the reticuloendothelial system (RES) (Brigger, Dubernet, and Couvreur<br />

2002). Nanoparticles coated with hydrophilic polymers have prolonged half-lives, believed to<br />

result from decreased opsonization and subsequent clearance <strong>by</strong> macrophages (Moghimi and<br />

Szebeni 2003).<br />

Cancer therapy has benefited from the use of liposomal doxorubicin, a formulation that again<br />

increases the therapeutic index of the active agent through a combination of passive tumor targeting<br />

and reduced toxicity (Gabizon and Martin 1997). In this case, coating the liposome with PEG<br />

significantly decreases uptake <strong>by</strong> macrophages and allows the liposomes to concentrate in tumors<br />

<strong>by</strong> escaping from the leaky vasculature surrounding solid tumors (Dvorak et al. 1988) through EPR<br />

effect (Maeda 2001). The surface chemistry of the nanoshells can be modified with PEG to increase<br />

biocompatibility and with sulfide-based linkers to allow the particles to be functionalized with<br />

targeting ligands.<br />

27.5 DENDRIMER AS DRUG DELIVERY SYSTEM: SELECTION,<br />

APPLICABILITY, AND RATIONALITY<br />

Dendrimers are a relatively new class of polymers with structures that depart rather dramatically<br />

from traditional linear polymers that are built from so-called AB monomers (Newkome, Moorefield,<br />

and Vogtle 1996). Because they are built from ABn, monomers (where n is usually 2 or 3),<br />

dendrimers are highly branched and have three distinct structural features: a core, multiple peripheral<br />

(end-) groups, and branching units that link the two. The peripheral groups and branching units<br />

are together called dendrons, or informally, wedges. Branching units used to date have contained<br />

virtually every type of functional group, ranging from those comprising pure hydrocarbons or<br />

aromatic groups to modified carbohydrates and nucleic acids. Dendrimers are iteratively synthesized<br />

so that the number of intervening branching units between the core and one end-group<br />

(i.e., the number of layers), called the generation number, is determined <strong>by</strong> the number of synthetic<br />

cycles. In the commercially available poly(amidoamine) (PAMAM) dendrimers, ammonia represents<br />

the zeroth generation (Tomalia, Naylor, and Goddard 1990). As the generation number<br />

increases from one to seven, the number of peripheral groups follows the geometric series 3, 6,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


530<br />

Nanotechnology for Cancer Therapy<br />

12,.,192. These peripheral groups largely control the solubility of the compound so that even with<br />

highly hydrophobic internal units, a dendrimer will dissolve in water if the end groups are sufficiently<br />

hydrophilic. As will be discussed here, this feature and the high local concentration of end<br />

groups enable a number of applications in bio-organic chemistry. Dendrimers represent a new class<br />

of highly branched polymers whose interior cavities and multiple peripheral groups facilitate<br />

potential applications in biomedicine and bio-organic chemistry (Kim and Zimmerman 1998).<br />

Major advances in the past years were made in the synthesis and study of new carbohydrate,<br />

nucleic acid, and peptide dendrimers as well as in the use of dendrimers as magnetic resonance<br />

imaging contrast agents, agents for cellular delivery of nucleic acids, and scaffolds for<br />

biomimetic systems.<br />

Researchers are now trying to synthesize dendrimer clusters for targeted therapy. Scientists at<br />

the University of Michigan have devised a method to easily create multipurpose molecules for use<br />

in targeted anti-cancer therapy (NCI 2005). By attaching single-stranded DNA linkers to dendrimers,<br />

the researchers could bridge together individual dendrimer subunits with specific functions<br />

and create a multifunctional dendrimer cluster. This bridging technique could potentially lead the<br />

way to developing customized therapies for individual patients. Dendrimers are promising candidates<br />

for use as the base of anti-cancer drugs to which functional molecules can be attached.<br />

Ideally, an anti-cancer therapeutic agent would have multiple functional groups such as a<br />

targeting molecule to bind a cell, a radioactive compound to kill the cell, and a fluorescent<br />

probe so the process can be imaged. However, there are problems synthesizing these compounds,<br />

and different drugs would have to be designed for each different tumor with this approach.<br />

Dr. James Baker, Jr. and colleagues improved compound synthesis through a building block<br />

approach in one NCI-funded study. They created two single-function dendrimers; one with a<br />

molecule to bind folate receptors and the other with a fluorescein molecule for imaging. They<br />

then added complimentary 34-base stretches of single-stranded DNA to each, and the two dendrimers<br />

conjugated through base pairing. Using fluorescence imaging, they observed that the<br />

dendrimer cluster specifically bound to a cancer cell line over expressing the folate receptor.<br />

Compared with traditional chemistry, DNA-linked dendrimers could provide a more effective<br />

way to mix and match different therapeutic combinations to better treat individual patients.<br />

Dendrimers represent one nanostructured material that may soon find its way into medical<br />

therapeutics. Starburst dendrimers are tree-shaped synthetic molecules with a regular branching<br />

structure emanating outward from a core that forms nanometer <strong>by</strong> nanometer with the number of<br />

synthetic steps or generations dictating the exact size of the particles, typically a few nanometers in<br />

spheroidal diameter (Freitas 2005). The peripheral layer can be made to form a dense field of<br />

molecular groups that serve as hooks for attaching other useful molecules such as DNA that can<br />

enter cells while avoiding triggering an immune response unlike viral vectors commonly employed<br />

today for transfection. Upon encountering a living cell, dendrimers of a certain size trigger a<br />

process called endocytosis where the cell’s outermost membrane deforms into a tiny bubble or<br />

vesicle. The vesicle encloses the dendrimer that is then admitted into the cell’s interior. Once inside,<br />

the DNA is released and migrates to the nucleus where it becomes part of the cell’s genome. The<br />

technique has been tested on a variety of mammalian cell types (Kukowska-Latallo et al. 2000) and<br />

in animal models (Vincent et al. 2003) though clinical human trials of dendrimer gene therapy<br />

remain to be done.<br />

Glycodendrimer nanodecoys have also been used to trap and deactivate some strains of influenza<br />

virus particles (Reuter et al. 1999; Landers et al. 2002). James Baker’s group at the University<br />

of Michigan is extending this work to the synthesis of multi-component nanodevices called tectodendrimers<br />

built up from a number of single-molecule dendrimer components (http://nano.med.umich.edu/projects/Dendrimers.html;<br />

Quintana et al. 2002). Tecto-dendrimers have a single core<br />

dendrimer surrounded <strong>by</strong> additional dendrimer modules of different types, each type designed to<br />

perform a function necessary to a smart therapeutic nanodevice (Figure 27.2). Baker’s group has<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 531<br />

Targeting<br />

Molecule<br />

Water<br />

10 –1<br />

Liposome<br />

Corona<br />

Glucose Antibody Virus Bacteria Cancer Cell A Period Tennis Ball<br />

Core<br />

Image Contrast<br />

Agent<br />

1 10 10 2<br />

10 3<br />

built a library of dendrimeric components from which a combinatorially large number of nanodevices<br />

can be synthesized.<br />

The initial library contains components that will perform the following tasks: diseased cell<br />

recognition, diagnosis of disease state, drug delivery, report location, and report outcome of<br />

therapy. By using this modular architecture, an array of smart therapeutic nanodevices can be<br />

created with little effort. For instance, once apoptosis-reporting, contrast-enhancing, and chemotherapeutic-releasing<br />

dendrimer modules are made and attached to the core dendrimer, it should be<br />

possible to make large quantities of this tecto-dendrimer as a starting material. This framework<br />

structure can be customized to fight a particular cancer simply <strong>by</strong> substituting any one of many<br />

possible distinct cancer recognition or targeting dendrimers, creating a nanodevice customized to<br />

destroy a specific cancer type and no other while also sparing the healthy normal cells. These<br />

nanodevices were synthesized using an ethylenediamine core with folic acid, fluorescein, and<br />

methotrexate covalently attached to the surface to provide targeting, imaging, and intracellular<br />

drug delivery capabilities (Freitas 1999). The “targeted delivery improved the cytotoxic response of<br />

the cells to methotrexate 100-fold over free drug” (Quintana et al. 2002).<br />

At least a half dozen cancer cell types have already been associated with at least one unique<br />

protein that targeting dendrimers could use to identify the cell as cancerous, and as the<br />

10 4<br />

Dendrimer Gold Nanoshell<br />

Drugs<br />

PEG<br />

(polyethylene<br />

glycol)<br />

Research<br />

Drug Screening<br />

(Labeling)<br />

Gene Delivery<br />

(Transfection)<br />

Diagnosis<br />

(Devices and<br />

Labeling)<br />

10 5<br />

10 6<br />

Quantum Dot<br />

Nanotechnology<br />

10 7<br />

Medical Applications<br />

10 8<br />

Nanometers<br />

Fullerene<br />

Clinical<br />

Drug Delivery<br />

(Therapy)<br />

Detection<br />

(Imaging)<br />

Diagnosis/Monitoring<br />

(Disease Markers)<br />

FIGURE 27.2 Morphological and size-based comparison of various novel drug delivery carriers such as<br />

liposome, dendrimer, gold nanoshell, quantum dot, fullerene, and nanometric scale of representation of<br />

various similar systems. (Adapted from McNeil, S. E., J. Leukoc. Biol., 78, 1–10, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


532<br />

genomic revolution progresses, it is likely that proteins unique to each kind of cancer will<br />

be identified, allowing Baker to design a recognition dendrimer for each type of cancer (http://nano.med.umich.edu/projects/Dendrimers.html).<br />

The same cell-surface protein recognition-targeting<br />

strategy could be applied against virus-infected cells and parasites. Molecular modeling has been<br />

used to determine optimal dendrimer surface modifications for the function of tecto-dendrimer<br />

nanodevices and to suggest surface modifications that improve targeting (Quintana et al. 2002).<br />

NASA and the National Cancer Institute have funded Baker’s lab to produce dendrimer-based<br />

nanodevices that can detect and report cellular damage because of radiation exposure in astronauts<br />

on long-term space missions (Sparks 2002). By mid-2002, the lab had built a dendrimeric nanodevice<br />

to detect and report the intracellular presence of caspase-3, one of the first enzymes released<br />

during cellular suicide or apoptosis (programmed cell death) that is one sign of a radiation-damaged<br />

cell. The device includes one component that identifies the dendrimer as a blood sugar so that the<br />

nanodevice is readily absorbed into a white blood cell and a second component using fluorescence<br />

resonance energy transfer (FRET) that employs two closely bonded molecules. Before apoptosis,<br />

the FRET system stays bound together, and the white cell interior remains dark upon illumination.<br />

Once apoptosis begins and caspase-3 is released, the bond is quickly broken, and the white blood<br />

cell is awash in fluorescent light. If a retinal scanning device measuring the level of fluorescence<br />

inside an astronaut’s body reads above a certain baseline, counteracting drugs can be taken.<br />

27.5.1 DIFFERENCES BETWEEN HYPERBRANCHED STRUCTURES AND DENDRIMERS<br />

Dagani (1996) discussed and compared hyperbranched polymers and dendrimers. Hyperbranched<br />

polymers have highly branched structure, but unlike dendrimers, their structure is neither regular<br />

nor highly symmetrical. The dendrimers were obtained <strong>by</strong> careful stepwise growth of successive<br />

layers or generations. Hyperbranched polymers are obtained in a single step <strong>by</strong> polycondensation of<br />

A 2B monomers that contain two reactive groups of type A and one of type B. Functional groups A<br />

and B are selected in a way that they react with each other to form a covalent bond. Structure of<br />

hyperbranched polymers is thought to be intermediate between those of linear polymers and<br />

dendrimers. Because of steric constraints and statistical nature of coupling steps, hyperbranched<br />

polymers have irregular structures as some of the A groups remain unreacted, leading to incorporation<br />

of linear segments (i.e., monomers units coupled at two rather than three points) (Frechet<br />

1994). Voit, Turner, and Mourey (1993) showed that viscosity-MW behavior of the hyperbranched<br />

polymers does not show a maximum as observed in the case of dendrimers. Instead, they obey the<br />

Mark–Hownick–Sakunada relation. Although their viscosities are anomalously lower when<br />

compared to linear polymers of comparable size. Analogous to dendrimers, A 2B hyperbranched<br />

polymers had essentially same number of reactive groups as they have monomeric units. However,<br />

unlike dendrimers, these functional groups are not all located at chain ends (Frechet 1994).<br />

27.5.2 CLASSIFICATION OF DENDRIMERS<br />

Nanotechnology for Cancer Therapy<br />

The dendrimers that are most studied and reported up to higher generations (Bosman, Janssen, and<br />

Meijer 1999) can be classified chemically viz. peptide dendrimers that were patented and described<br />

upto higher generations <strong>by</strong> Denkewalter in the early 1980s. Those can well be characterized <strong>by</strong><br />

size-exclusion chromatography only. PAMAM dendrimers are another class of dendritic structures<br />

that have been thoroughly investigated and had received wide spread attention. These were divergently<br />

constructed, e.g., PAMAM described <strong>by</strong> Tomalia and arborol systems of Newkome.<br />

Polypropyleneimine dendrimers were produced <strong>by</strong> Vogtle <strong>by</strong> divergent synthesis. It was also<br />

produced <strong>by</strong> heterogeneously catalyzed hydrogenation <strong>by</strong> De Brabander-van-den Berg and<br />

Meijer (1993) and is also studied widely. Frechet produced Poly-ether dendrimers <strong>by</strong> convergent<br />

procedures (Wooley et al. 1994). These also include Poly(aryl ether) dendrimers. Moore produced<br />

Phenyl acetylene dendrimers <strong>by</strong> convergent approach. Tetrathiofulvalene (TTF) 21-glycol<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 533<br />

dendrimers are another class of dendrimers that are intra-dendrimer aggregates of TTF cation<br />

radicals (Christensen et al. 1998). These are evident <strong>by</strong> spectrochemical studies of partially<br />

oxidized TTF units. These were prepared <strong>by</strong> convergent approaches. Pentaporphyrin is another<br />

starburst porphyrin polymer that is considered as a first generation dendrimers (Norsten and Branda<br />

1998). These had hybrid properties of porphyrin and dendrimers where ether linkages are found for<br />

iterations. Aryl ester monodispersed dendrimers based on 1, 3, 5-benzene tricarboxylic acid were<br />

initially described <strong>by</strong> Miller, Kwock, and Neenan (1992) and prepared <strong>by</strong> convergent synthesis.<br />

This was possibly used for MWs standards as polymeric rheological modifier, molecular inclusion<br />

hosts, etc. Two types of dendrimeric nucleic acids have been studied, viz. those that are covalently<br />

connected, and those that self-assemble via complementary base-pairing schemes (Kim and<br />

Zimmerman 1998). An example of a covalent DNA dendrimer was reported, wherein a secondgeneration<br />

phosphodiester-based dendrimer derived from pentaerythritol held nine pentathymidylate<br />

units on the end-groups and either a 15-mer or 26-mer at the core.<br />

Additionally, many other types of interesting, valuable, aesthetically pleasing dendritic systems<br />

have been developed such as Dendrophanes that were first described <strong>by</strong> Diederich and coworkers.<br />

They called these phenyl methane-based cyclophane. These were designed as globular proteins.<br />

Metallodendrimers include Ruthenium-terpyridine complexed dendrimers <strong>by</strong> Newkome and<br />

coworkers, dendritic iron (II) complexes <strong>by</strong> Chow and coworkers, zinc-porphyrin dendrimers <strong>by</strong><br />

Diederich and coworkers, etc. and they contain metal complexed in dendrimer structure. Polyamino<br />

phosphine containing dendrimers are phosphorous containing dendrimers. Mesogen functionalized<br />

carbosilane dendrimer involves functionalization <strong>by</strong> 36 mesogenic units attached through C-5<br />

spacer to liquid crystalline dendrimer that form smectic-A phase in a temperature range of 117–<br />

1308C. Dendritic box was based on construction of a chiral shell of protected amino acids onto<br />

polypropyleneimine dendrimers with 64 amino end groups. These monodispersed dendritic<br />

containers of nanometric dimensions have physically entrapped or locked in guest molecules<br />

(Jansen et al. 1994; Jansen and Meijer 1995). Toyokuni et al. (1994) described an example of<br />

carbohydrate dendrimers for tumor-associated carbohydrate antigens. They linked it without the<br />

use of macromolecular carrier or an adjuvant, and they conjugated it with starburst PAMAM<br />

dendrimers to elicit antibody responses.<br />

Dendrimeric systems can also be classified on the basis of physical characteristics of the system<br />

such as simple dendrimers having simple monomeric units; liquid crystalline dendrimers having<br />

mesogenic liquid crystalline substances; chiral dendrimers having optical activities; and micellar<br />

dendrimers having unimolecular micellar structures or water-soluble hyperbranched dendrimers of<br />

polyphenylene, etc. Hybrid dendrimers are combinations of dendritic and linear polymers in hybrid<br />

block or graft copolymers forms, whereas amphiphilic dendrimers are a class of globular dendrimers<br />

having unsymmetrical, but highly controlled, distributions of chain end chemistry. These are<br />

oriented at interfaces forming interfacial liquid membranes for stabilizing aqueous organic emulsion.<br />

These may be oriented under influence of external stimulus, e.g., electric field.<br />

27.5.3 FEATURES OF DENDRIMERS AS NANODRUGS<br />

The various features of dendrimeric formulations in use include cost and ease of manufacture for<br />

cost effective (compared with small molecule drugs), and such formulations can be scaled-up for<br />

Current Good Manufacturing Practices (cGMP) guidelines manufacture. These are effective and<br />

active against a wide range of diseases and for their novel modes of action. Also at therapeutic<br />

doses, the dendrimers are safe and stable as solids, and they are thus available in a variety of<br />

Pharmaceutical formulations. The formulations are reproducible and defined as far as identity is<br />

concerned. They appear in most cases as white to off-white powder. Starpharma-like Pharmaceuticals<br />

(www.starpharma.com/framemaster.htm) are creating value from such dendrimer-based<br />

nanotechnology. Their core business strategies include developing high-value dendrimer nanodrugs<br />

to address unmet market needs, partnering with pharmaceutical companies to create new<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


534<br />

opportunities and solutions to problems with the application of dendrimer nanotechnology,<br />

extending core skills and know-how through licensing and partnering with others, and investing<br />

in non-pharmaceutical applications of dendrimer technology. Rational drug design is being used <strong>by</strong><br />

the biotechnology and pharmaceutical industry to design novel drugs based on known biological<br />

targets (e.g., receptors involved in the pathogenesis of disease). In this context, Starpharma’s<br />

dendrimer technologies offer a unique ability to design functionalized, polyvalent dendrimers as<br />

defined species for specific biological targets.<br />

The dendrimers are being eyed as drug-delivery agents, micelle mimics, and nanoscale building<br />

blocks. There are also added possibilities for grafting of active groups on surface of dendrimer <strong>by</strong><br />

using cross-linkers. Also fatty acids, and polymers such as poly(ethyleneglycol) (PEG), etc. can be<br />

attached to dendrimers due to –COOH and –NH 2 terminal active groups present on its surface. PEG<br />

coatings, also called as PEGylation, could also have many possible numerous advantages so could<br />

be selected for coating (Zeng and Zimmerman 1997). Gitsov and Fréchet (1996) published some<br />

preliminary results on macromolecules that change shape when the polarity of the solvent changes.<br />

These macromolecules consist of four long PEG chains extending out from a central carbon atom.<br />

Each of these hydrophilic PEG arms is terminated with a hydrophobic wedge-like dendritic group<br />

based on 3,5-dihydroxybenzyl alcohol. The arms are long enough and flexible enough to allow this<br />

hybrid star to assume several very different conformations in solution. In tetrahydrofuran, the<br />

hybrid star forms a unimolecular micelle that has a hydrophilic core of tightly packed PEG arms<br />

surrounded <strong>by</strong> a loose hydrophobic shell of dendritic wedges. In chlorinated solvents such as<br />

chloroform where both building blocks are readily soluble, the PEG arms (and their dendritic<br />

extremities) extend outward, making the core more accessible. In polar or aqueous media such<br />

as methanol, the hydrophilic PEG arms loop around the hydrophobic dendritic wedges, pushing the<br />

wedges into the star’s core and leaving an outer PEG layer exposed to the outside world. The hybrid<br />

stars rearrange their two types of components in this way to minimize the overall free energy of<br />

the system.<br />

A novel hybrid star prepared in Fréchet’s lab at the University of California Berkeley changes<br />

its conformation when the polarity of the solvent medium changes. In THF (tetrahydrofuran), the<br />

hydrophilic PEG core is compact, whereas in chloroform, it is more extended and open. In<br />

methanol, the hydrophilic PEG arms envelope the hydrophobic dendritic wedges, leading to two<br />

possible conformations. In one, the dendritic wedges are tightly packed toward the middle of the<br />

micelle with the PEG arms forming loops in the surrounding medium. In the other arrangement,<br />

each dendritic moiety is individually wrapped in a PEG arm and somewhat extended into the<br />

medium. Freemantle (1999) also explored blossoming of dendrimers in a conference that explored<br />

design, synthesis, structure, and potential applications of highly branched macromolecules.<br />

27.6 APPLICATIONS OF PEGYLATED DENDRIMERS<br />

Nanotechnology for Cancer Therapy<br />

Dendrimers are synthetic, nanoscale structures (1–100 nm) types of polyvalent pharmaceuticals<br />

(www.starpharma.com/framemaster.htm) that can be tailored for many pharmaceutical applications.<br />

Specialized chemistry techniques allow for precise control of the physical and chemical<br />

properties of dendrimers. They are constructed in a series of controlled steps (see below) that<br />

increase the number of small branching molecules around a central core molecule. The final<br />

generation of molecules added to the growing structure makes up the polyvalent surface of the<br />

dendrimer. The core branching and surface molecules are chosen to give desired properties and<br />

functions. Dendrimers allow researchers, for the first time, to produce highly defined and biocompatible<br />

nanoscale objects built from the bottom up. This high definition enables unique<br />

functionality in life sciences applications. For readers interested in a historical account, terminology,<br />

nomenclature, or other background material, an excellent monograph <strong>by</strong> Newkome,<br />

Moorefield, and Vogtle (1996) is recommended. Zeng and Zimmerman (1997) broadly reviewed<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 535<br />

applications of dendrimers in bioorganic chemistry that have been reported in the past year, and,<br />

although somewhat different in focus, it updates earlier reviews published. An excellent review of<br />

dendrimers that act as biological mimics was also published <strong>by</strong> Smith and Diederich (1998).<br />

Beyond discussing a couple of results in the area of biomimetic dendrimers, this review highlighted<br />

various work on dendrimers whose monomer units are derived from carbohydrates, amino acids, or<br />

nucleotides and also several promising biomedical applications under active investigation.<br />

27.6.1 ENHANCED BIOCOMPATIBILITY FOR DRUG DELIVERY<br />

PAMAM dendrimers having poly(ethylene glycol) (PEG) grafts were designed as a novel drug<br />

carrier that possesses an interior for the encapsulation of drugs and a biocompatible surface. PEG<br />

monomethyl ether with the average MW of 550 or 2000 Da was combined to essentially every<br />

chain end of the dendrimer of the third or fourth generation via urethane bond (Kojima et al. 2000).<br />

The PEG-attached dendrimers encapsulating anti-cancer drugs adriamycin and methotrexate were<br />

prepared <strong>by</strong> extraction with chloroform from mixtures of the PEG-attached dendrimers and varying<br />

amounts of the drugs. Their ability to encapsulate these drugs increased with increasing dendrimer<br />

generation and chain length of PEG grafts. Among the PEG-attached dendrimers prepared, the<br />

highest ability was achieved <strong>by</strong> the dendrimer of the fourth generation having the PEG grafts with<br />

the average MW of 2,000 Da that could retain 6.5 adriamycin molecules or 26 methotrexate<br />

molecules/dendrimer molecule. The methotrexate-loaded PEG-attached dendrimers slowly<br />

released the drug in an aqueous solution of low ionic strength. However, in isotonic solutions,<br />

methotrexate and doxorubicin were readily released from the PEG-attached dendrimers.<br />

Padilla et al. (2002) described the use of high MW polymers (O 20,000 Da) as soluble drug<br />

carriers to improve drug targeting and therapeutic efficacy. They evaluated various dendritic architectures<br />

composed of a polyester dendritic scaffold based on the monomer unit 2,2bis(hydroxymethyl)<br />

propanoic acid for their suitability as drug carriers both in vitro and in vivo.<br />

These systems are both water-soluble and non-toxic. In addition, the potent anti-cancer drug<br />

doxorubicin was covalently bound via a hydrazone linkage to a high MW 3-arm poly(ethylene<br />

oxide) (PEO)-dendrimer hybrid. Drug release was a function of pH, and the release rate was more<br />

rapid at pH ! 6. The cytotoxicity of the DOX-polymer conjugate measured on multiple cancer<br />

lines in vitro was reduced but not eliminated, indicating that some active doxorubicin was released<br />

from the drug polymer conjugate under physiological conditions. Furthermore, biodistribution<br />

experiments show little accumulation of the DOX-polymer conjugate in vital organs, and the<br />

serum half-life of doxorubicin attached to an appropriate high MW polymer has been significantly<br />

increased when compared to the free drug. Therefore, this new macromolecular system exhibits<br />

promising characteristics for the development of new polymeric drug carriers.<br />

Kolhe et al. (2003) attached PEG grafts to the terminal amine groups of the dendrimer and<br />

enhanced biocompatibility of higher generation PAMAM dendrimer. Similarly, Bhadra et al.<br />

(2003) described the use of uncoated and PEGylated newer PAMAM dendrimers for delivery of<br />

an anti-cancer drug 5-fluorouracil. The 4.0 G PAMAM dendrimer was PEGylated using N-hydroxysuccinimide-activated<br />

carboxymethyl MPEG-5000. The PEGylation of the systems was found to<br />

have increased their drug-loading capacity, reduced their drug release rate, and hemolytic toxicity.<br />

TEM study revealed quite uniform surface of the systems. The systems were found suitable for<br />

prolonged delivery of an anti-cancer drug <strong>by</strong> in vitro and blood-level studies in albino rats without<br />

producing any significant hematological disturbances. PEG modification has been found to be<br />

suitable for modification of PAMAM dendrimers for reduction of drug leakage and hemolytic<br />

toxicity. This, in turn, could improve drug-loading capacity and stabilize such systems in the<br />

body. The study suggests use of such PEGylated dendrimeric systems as nanoparticulate depot<br />

type of system for drug administration (Figure 27.3).<br />

The entrapment of 5-fluorouracil in PEG modification dendrimers significantly increased <strong>by</strong> 12<br />

times because of more sealing of dendrimeric structure <strong>by</strong> PEG coating at the peripheral portions of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


536<br />

FIGURE 27.3 Chemical structural representation of PEG coated PAMAM loaded with 5-fluorouracil.<br />

(Adapted from Bhadra, et al., Int. J. Pharm., 257, 111, 2003. With permission.)<br />

dendrimers as coat that prevented drug release <strong>by</strong> enhancing complexation probably <strong>by</strong> steric and<br />

electronic effects of the additional functional groups made available <strong>by</strong> MPEG (Figure 27.4). This,<br />

in turn, also reduced the rate of drug release from such systems across dialysis membrane as<br />

observed <strong>by</strong> their release profile. The PAMAM dendrimers on PEGylation behaved similar to<br />

Cumulative % drug release<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

4.0 G non-PE Gylated 4.0G PEGylated<br />

1 2 3 4 5 6 7 24 48 72 96 120 144<br />

Time interval (hr.)<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 27.4 Comparative cumulative percentage release of 5-fluorouracil from PEG coated and uncoated<br />

PAMAM dendrimers. (Adapted from Bhadra, et al., Int. J. Pharm., 257, 111, 2003. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 537<br />

(Concentration in blood (μg/ml)<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

0 100 200 300 400 500 600 700 800<br />

Time (min)<br />

spherical unimolecular polymer micelles with chains of hydrated methoxy-poly(ethyleneglycol) on<br />

its surface as coat. This increased the entrapment capacity of dendrimers and made them act as<br />

nanometric containers carrying drugs. Such PEGylated dendrimers can be suggested for sustaining<br />

delivery of the drug 5-fluorouracil as also observed <strong>by</strong> the blood level data where blood level was<br />

much prolonged and was detectable up to 12 h. The formulations were found to be following<br />

sustained release characteristics for 5-FU (Figure 27.5) as shown <strong>by</strong> the relative increase in<br />

MRT for both non-PEGylated and PEGylated drug-dendrimer complexes as compared to plain<br />

drug solution (6.024 and 13.31 times, respectively). The release rate, however, was found to have<br />

increased in vivo as compared to in vitro data, possibly because of the metabolism <strong>by</strong> the enzymes<br />

and hydrolysis in the body. The blood level of the drug was found to be lower in case of PEGylated<br />

systems than that of non-PEGylated as a result of slower release rates of the drug similar to the trend<br />

found in vitro for the drug release. This might also be attributed to better cellular penetration and<br />

adhesion properties of the PEG chains of PEGylated systems that might lead to entanglement of<br />

such systems within capillary fenestration from which such systems might release drugs acting as<br />

nanoparticulate depot type carrier present in blood circulation.<br />

27.6.2 MICELLAR MEANS OF ENHANCING SOLUBILITY<br />

5-FU-PAMAM dend (non-PE Gylated)<br />

5-FU-PAMAM dend (PE Gylated)<br />

FIGURE 27.5 Blood level data of 5-fluorouracil from PEG coated and uncoated PAMAM dendrimers.<br />

(Adapted from Bhadra, et al., Int. J. Pharm., 257, 111, 2003. With permission.)<br />

Chapman and coworkers prepared amphiphilic copolymers derived from linear PEO and tBoCterminated<br />

poly-a,2-lysine dendrimers. The use of PEO as a platform for the dendrimers synthesis<br />

greatly facilitated separation because product up to fourth generation could be precipitated from<br />

reaction mixtures <strong>by</strong> ether. Surface tension method was employed to determine Critical Micelle<br />

Concentration (CMC) and micelle formation behavior. The surface of the aggregates was suggested<br />

to be highly compact. The existence of transition concentration for dye solubility of dye orange OT<br />

<strong>by</strong> G4 hydramphiphiles in water supports the micellar behavior (Chapman et al. 1994).<br />

Frechet and coworkers also synthesized another novel class of amphiphilic star copolymers. A<br />

four-armed PEG star was attached to four polyether dendrons derived from a pentaerythritol core.<br />

Results of SEC/VISC and 1 HNMR studies indicated the formation of unimolecular micelles in<br />

chloroform, tetrahydrofuran, or methanol, but with strikingly different structures. The star copolymers<br />

could self organizes into different micellar structures as a function of the environment. A<br />

potential application of these stimuli responsive copolymers involves their uses as solvent specific<br />

encapsulation agents (Gitsov and Frechet 1996).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


538<br />

Water-soluble dendritic unimolecular micelles as potential drug delivery agents had been<br />

explored <strong>by</strong> Liu et al. (2000) with the hydrophobic core surrounded <strong>by</strong> hydrophilic shell. These<br />

were prepared <strong>by</strong> coupling dendritic hypercores with PEG-mesylates. The monomeric core that was<br />

selected to build it was 4,4-bis(4 0 hydroxy phenyl) pentanol as this larger monomeric unit provides<br />

flexibility to the dendritic structures while contributing to the container capacity of the overall<br />

structure. Four generations of dendritic hypercores with 6, 12, 24, and 48 phenolic end groups were<br />

prepared. Subsequent coupling reaction with PEG-mesylates afforded four generations of dendritic<br />

unimolecular micelles (Gitsov and Fréchet 1993). The container property was demonstrated <strong>by</strong><br />

pyrene solubilization in aqueous solution. In general, copolymers containing low generation<br />

dendrons, especially G1, tended to form unimolecular micelle, whereas G2 and G3 copolymers<br />

formed multi molecular micelles, presumably driven <strong>by</strong> hydrophobic effects and p–p interaction<br />

between dendritic blocks. Coexistence of two well-separated peaks in the SEC indicated a slow<br />

exchange process between these two species of different sizes.<br />

Ooya, Lee, and Park (2003) developed new methods and pharmaceutical compositions to<br />

increase the aqueous solubility of paclitaxel (PTX), a poorly water-soluble drug. Graft and starshaped<br />

graft polymers consisting of PEG400 graft chains increased the PTX solubility in water <strong>by</strong><br />

three orders of magnitude. Polyglycerol dendrimers (dendriPGs) dissolved in water at high concentrations<br />

without significantly increasing the viscosity and at 80 wt% were found to increase the<br />

solubility of PTX 10,000-fold. The solubilized PTX was released from graft polymers, star-shaped<br />

graft polymers, and the dendriPGs into the surrounding aqueous solution. The release rate was a<br />

function of the star shape and the dendrimer generation. The availability of the new graft, star, and<br />

dendritic polymers having ethylene glycol units should permit development of novel delivery<br />

systems for other poorly water-soluble drugs.<br />

27.6.3 ENHANCING PHARMACOKINETIC PROPERTIES<br />

Nanotechnology for Cancer Therapy<br />

New polyester dendrimer-PEO bow-tie hybrids were evaluated for their potential as drug delivery<br />

vehicles and to explore the effect of MW and architecture on their pharmacokinetic properties<br />

(Elizabeth et al. 2005). In vitro experiments showed that these polymers were non-toxic to MDA-<br />

MB-231 cells and that they degraded to lower MWs at the normal physiological pH of 7.4 and at the<br />

mildly acidic pH of 5.0 that may be encountered upon uptake of these molecules <strong>by</strong> endocytosis and<br />

subsequent trafficking to endosomes and lysosomes. Biodistribution studies showed that all carriers<br />

with MWs of 40,000 and greater had long plasma circulation times, whereas those with lower MWs<br />

were cleared more rapidly with significant quantities excreted in the urine. In general, it was found<br />

that the more branched [G-3] polymers exhibited increased circulation times and decreased renal<br />

clearance relative to the less branched polymers, a property likely attributable to their decreased<br />

flexibility and resulting difficulty in passing through glomerular pores. It was also demonstrated that<br />

the more branched structures provide increased levels of steric protection for the payload on the<br />

drug-carrying dendron at the core of the molecule. High levels of tumor accumulation were found<br />

for two [G-3] polymers in mice bearing subcutaneous B16F10 melanoma. Overall, the features of<br />

these new branched carriers include degradability, lack of toxicity, long circulation half-lives, and<br />

high levels of tumor accumulation make them very promising polymers for therapeutic applications.<br />

High MW polymers have shown promise in terms of improving the properties and the<br />

efficacy of low MW therapeutics. However, new systems that are highly biocompatible are biodegradable,<br />

have well-defined MW, and have multiple functional groups for drug attachment are still<br />

needed. The biological evaluation of a library of eight polyester dendrimer-PEO bow-tie hybrids is<br />

described here.<br />

The group of evaluated polymers was designed to include a range of MWs (from 20,000 to<br />

160,000) and architectures with the number of PEO arms ranging from two to eight. In vitro<br />

experiments revealed that the polymers were non-toxic to cells and were degraded to lower MW<br />

species at pH 7.4 and pH 5.0. Biodistribution studies with 125 I-radiolabeled polymers showed that<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 539<br />

the high MW carriers (O40,000) exhibited long circulation half-lives. Comparison of the renal<br />

clearances for the four-arm versus eight-arm polymers indicated that the more branched polymers<br />

were excreted more slowly into the urine, a result attributed to their decreased flexibility. Because<br />

of their essentially linear architecture that does not provide for good isolation of the iodinated<br />

phenolic moieties, the polymers with two arms were rapidly taken up <strong>by</strong> the liver. The biodistributions<br />

of two long-circulating high MW polymers in mice bearing subcutaneous B16F10 tumors<br />

were evaluated, and high levels of tumor accumulation were observed. These new carriers are<br />

promising for applications in drug delivery and are also useful for improving the understanding of<br />

the effect of polymer architecture on pharmacokinetic properties (Elizabeth et al. 2005).<br />

27.6.4 DENDRIMER USE AS TRANSFECTION REAGENTS<br />

The recent completion of the human genome sequence has provided promising tools for defining<br />

cancer-specific targets, including genes and their expression patterns. In addition, the development<br />

in proteomics and achievements of high-throughput screenings offer great potential for discovering<br />

effective drugs, including DNA drugs, for gene therapy. Therefore, cancer therapy is no longer<br />

limited <strong>by</strong> the identification of genes; it is limited <strong>by</strong> an inability to deliver them efficiently to<br />

cancer cells, and the bottleneck in cancer gene therapy will soon be shifting from gene discovery<br />

to gene delivery (Luo 2004). The goal is to provide adequate amounts of drugs that are close to<br />

targeted cells because the actual destination for all gene delivery is the cell nucleus. Only drugs that<br />

are close to cells can be internalized. Once inside, intracellular barriers are considerably different.<br />

For example, diffusion is no longer the bottleneck. Rather, cytosol survival and nuclear targeting<br />

are more important (for non-viral gene delivery). An important and difficult task for cancer gene<br />

therapy is increasing the expression level of the therapeutic gene (that is usually toxic) in targeted<br />

tumor cells, but at the same time, avoiding its expression in normal tissue (Luo and Saltzman 2000).<br />

Methoxypoly(ethylene glycol)-block-poly(L-lysine) dendrimer was designed to form a watersoluble<br />

complex with plasmid DNA. The copolymer was synthesized <strong>by</strong> the liquid-phase peptide<br />

synthesis method (Choi et al. 1999). Agarose gel electrophoresis and DNase I protection assay<br />

proved that this linear polymer/dendrimer block copolymer spontaneously assembled with plasmid<br />

DNA, forming a water-soluble complex that increased the stability of the complexed DNA. Atomic<br />

force microscopy of the complex was evaluated at various charge ratios, showing that the copolymer/DNA<br />

complex was like a globular shape.<br />

Choi et al. (2000) synthesized and used a barbell-like triblock copolymer, poly(L-lysine)<br />

dendrimer-block-poly(ethylene glycol)-block-poly(L-lysine) dendrimer, for delivery of plasmid<br />

DNA <strong>by</strong> its self-assembly with the structures. A barbell-like ABA-type triblock copolymer,<br />

poly(L-lysine) dendrimer-block-poly(ethylene glycol)-block-poly(L-lysine) dendrimer<br />

(PLLD–PEG–PLLD), was synthesized <strong>by</strong> the liquid-phase peptide synthesis method. The selfassembling<br />

complex formation of the third and fourth generation of the copolymer with plasmid<br />

DNA was studied. 1 H NMR and matrix-assisted laser desorption/ionization-time-of-flight mass<br />

spectrometry (MALDI-TOF MS) were used for the characterization of the synthesized copolymer.<br />

The self-assembling behavior of the third and fourth generations of the copolymer with plasmid<br />

DNA was investigated <strong>by</strong> electrophoretic mobility shift assay, DNase I protection assay, and<br />

ethidium bromide exclusion assay. They observed great differences in the self-assembling ability<br />

of the third and fourth generations of the polymer. This suggests that the number of positively<br />

charged amines per polymer molecule should be an important factor for the potential for selfassembling<br />

complex formation with DNA. Atomic force microscopy (AFM) and z potentials were<br />

used for evaluating the shape, size distribution, and surface charge of the complexes at various<br />

charge ratios. From AFM images, it was observed that the shape of the complex was nearly<br />

spherical, and its size was about 50–150 nm in diameter. The in vitro cytotoxicity of the copolymer<br />

was compared with that of poly(L-lysine), poly(D-lysine), and polyethylenimine.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


540<br />

Luo et al. (2002) demonstrated a simple and successful synthetic approach to devise a highly<br />

efficient DNA delivery system with low cytotoxicity and low cost. PAMAM dendrimer is a highly<br />

efficient DNA delivery agent when compared to other chemical transfection reagents. Partially<br />

degraded, high-generation dendrimers offer even higher efficiency, presumably because of<br />

enhanced flexibility of the otherwise rigid dendrimer chains. It was hypothesized that chemical<br />

modification of low generation dendrimer with biocompatible PEG chains would create a conjugate<br />

of PAMAM core with flexible PEG chains that mimics the fractured high-generation dendrimer and<br />

produces high transfection efficiency. Generation 5 PAMAM dendrimer was modified with<br />

3400 MW PEG. The novel conjugate produced a 20-fold increase in transfection efficiency<br />

compared with partially degraded dendrimer controls. The cytotoxicity of PEGylated dendrimers<br />

was very low. This extremely efficient, highly biocompatible, low-cost DNA delivery system can<br />

be readily used in basic research laboratories and may find future clinical applications.<br />

This system can be readily used in basic research laboratories and may be modified and applied<br />

in future clinical usage. Furthermore, similar approaches could be adopted with other DNA delivery<br />

agents that may create new opportunities for using DNA as a drug via a synthetic delivery route.<br />

Mannisto et al. (2002) investigated the influence of shape, MW, and PEGylation of linear,<br />

grafted, dendritic, and branched poly(L-lysine) on their DNA delivery properties. DNA binding,<br />

condensation, complex size and morphology, cell uptake, and transfection efficiency were<br />

determined. Most poly(L-lysine) condense DNA linear polymers and are at the same time more<br />

efficient than most dendrimers. At low MWs of PLL, DNA binding and condensation were less<br />

efficient, particularly with dendrimers. PEGylation did not decrease DNA condensation of PLLs at<br />

less than 60% (fraction of MW) of PEG. PEGylation sterically stabilized the complexes but did not<br />

protect them from interaction with polyanionic chondroitin sulfate. Cell uptake of poly(L-lysine)/<br />

DNA complexes was high and PEGylation increased the transfection efficacy. However, overall<br />

transfection level of poly(L-lysine) is low possibly as a result of the inadequate escape of the<br />

complexes from endosomes or poor release of DNA from the complexes. Physicochemical and<br />

biological structure-property relationships of poly(L-lysine) were demonstrated, but no clear correlations<br />

between the tested physicochemical determinants (size of complexes, zeta-potentials,<br />

condensation of DNA, and the shape of complexes) and biological activities were seen. Intracellular<br />

factors and/or still unknown features of DNA complexation may ultimately determine<br />

transfection activity with the carriers.<br />

Kim et al. (2004) synthesized and applied a novel triblock copolymer, PAMAM-block-PEGblock-PAMAM,<br />

as a gene carrier. PAMAM dendrimer is proven to be an efficient gene carrier<br />

itself, but it is associated with certain problems such as low water solubility and considerable<br />

cytotoxicity. Therefore, they introduced PEG to engineer a non-toxic and highly transfectionefficient<br />

polymeric gene carrier because PEG is known to convey water-solubility and biocompatibility<br />

to the conjugated copolymer. This copolymer could achieve self-assembly with plasmid<br />

DNA, forming compact nanosized particles with a narrow size distribution. Fulfilling expectations,<br />

the copolymer was found to form highly water-soluble polyplexes with plasmid DNA, showed little<br />

cytotoxicity despite its poor degradability, and finally achieved high transfection efficiency comparable<br />

to PEI. Consequently, they showed that an approach involving the introduction of PEG to<br />

create a tree-like cationic copolymer possesses a great potential for use in gene delivery systems.<br />

Logically, this approach could be extended to cancer treatment.<br />

27.6.5 STIMULI RESPONSIVE CARRIERS<br />

Nanotechnology for Cancer Therapy<br />

Gillies and Frechet (2002) designed and prepared a new polyester dendrimer, PEO hybrid systems,<br />

for drug delivery and related therapeutic applications. These systems consist of two covalently<br />

attached polyester dendrons where one dendron provides multiple functional handles for the attachment<br />

of therapeutically active moieties, whereas the other is used for attachment of solubilizing<br />

PEO chains. By varying the generation of the dendrons and the mass of the PEO chains, the MW,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 541<br />

architecture, and drug loading can be readily controlled. The bow-tie shaped dendritic scaffold was<br />

synthesized using both convergent and divergent methods with orthogonal protecting groups on the<br />

periphery of the two dendrons. PEO was then attached to the periphery of one dendron using an<br />

efficient coupling procedure. A small library of eight carriers with MWs ranging from about 20 to<br />

160 kDa were prepared and characterized <strong>by</strong> various techniques, confirming their welldefined<br />

structures.<br />

Gillies, Jonsson, and Frechet (2004) used a pH-responsive micelle system, linear-dendritic<br />

block copolymers comprising PEO and either a polylysine or polyester dendron. These were<br />

prepared, and hydrophobic groups were attached to the dendrimer periphery <strong>by</strong> highly acid-sensitive<br />

cyclic acetals. These copolymers were designed to form stable micelles in aqueous solution at<br />

neutral pH and disintegrate into unimers at mildly acidic pH following loss of the hydrophobic<br />

groups upon acetal hydrolysis. Micelle formation was demonstrated <strong>by</strong> encapsulation of the fluorescent<br />

probe Nile Red, and the micelle sizes were determined <strong>by</strong> dynamic light scattering. The<br />

structure of the dendrimer block, its generation, and the synthetic method for linking the acetal<br />

groups to its periphery all had an influence on the CMC and the micelle size. The rate of hydrolysis<br />

of the acetals at the micelle core was measured for each system at pH 7.4 and pH 5, and it was found<br />

that all systems were stable at neutral pH but that they underwent significant hydrolysis at pH 5 over<br />

several hours. The rate of hydrolysis at pH 5 was dependent on the structure of the copolymer, most<br />

notably the hydrophobicity of the core-forming block. To demonstrate the potential of these<br />

systems for controlled release, the release of Nile Red as a model payload was examined. At pH<br />

7.4, the fluorescence of micelle-encapsulated Nile Red was relatively constant, indicating it was<br />

retained in the micelle, and at pH 5, the fluorescence decreased, consistent with its release into the<br />

aqueous environment. The rate of release was strongly correlated with the rate of acetal hydrolysis<br />

and was controlled <strong>by</strong> the chemical structure of the copolymer. The mechanism of Nile Red release<br />

was investigated <strong>by</strong> monitoring the change in size of the micelles over time at acidic pH. Dynamic<br />

light scattering measurement showed a size decrease over time, eventually reaching the size of a<br />

unimer, providing evidence for the proposed micelle disintegration.<br />

27.6.6 ALTERATION OF TOXICITY<br />

In one of the studies Bhadra et al. (2003) found that PEGylation significantly decreased the<br />

hemolysis of red blood cells (RBCs) to below 5% in case of whole generation amine terminated<br />

charged dendrimers having about 15.3–17.3% hemolytic toxicity, similar to negligible in case of<br />

half generations of carboxylic acid terminated dendrimers. This was due to inhibition of interaction<br />

of RBCs with the charged quaternary ammonium ion as determined <strong>by</strong> interaction with RBCs. The<br />

hemolytic toxicity of the dendrimers was enough to preclude its use as a drug delivery system. The<br />

toxicity was due to the poly-cationic nature of the PAMAM dendrimers that was also responsible<br />

for their cytotoxicity (Malik et al. 2000), particularly in the case of whole generation amine<br />

terminated charged dendrimers, but not in the case of half generations of carboxylic acid<br />

terminated dendrimers.<br />

The PEGylation of the dendrimers were found to have also considerably lower rate of hemolysis<br />

of the RBCs because of significantly reduced interaction of RBCs with the charged quaternary<br />

ammonium ion that is generally present on the amine terminated whole generations of dendrimers.<br />

Also, the hematological study was undertaken to assess the relative effects of the PEGylated and<br />

non-PEGylated systems as compared to the plain drug on various blood parameters. The blood<br />

parameters undergoing major changes as to normal values of blood levels are RBC count, WBC<br />

count, and differential lymphocytes count. The RBC count of non-PEGylated 5-FU-dendrimers was<br />

found to have decreased below normal values <strong>by</strong> about 1!10 6 /ml RBCs to 2!10 6 /ml RBCs as<br />

against similar PEGylated systems. The WBC count of non-PEGylated 5-FU-dendrimer complex<br />

increased <strong>by</strong> 3!10 3 /ml cells to 4!10 3 /ml cells as compared to normal values. However, for 5<br />

FU-PEGylated dendrimer complexes, the increase was <strong>by</strong> 1!10 3 /ml cells to 1.5!10 3 /ml WBCs<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


542<br />

Nanotechnology for Cancer Therapy<br />

as compared to normal count in controlled group. The increase in WBC count is significant in case<br />

of non-PEGylated systems. Similarly, relatively greater increase in lymphocyte count was observed<br />

<strong>by</strong> non-PEGylated dendrimer-drug complexes that were <strong>by</strong> about 2!10 3 /ml cells. This was similar<br />

to blood toxicity and cytotoxicity effects of acrylates nanoparticulates (Malik et al. 2000), which are<br />

known to be stimulating the macrophage level and WBC count. This was found to be true in the<br />

case of these non-PEGylated drug-dendrimeric systems but not in case of PEGylated systems,<br />

conforming the trends of the PEGylated carriers (Veronese et al. 1989; Papahadjopoulus et al.<br />

1991; Phillips et al. 1999) undergoing lesser phagocytic uptake.<br />

Neerman et al. (2004) also showed that dendrimers based on melamine could reduce the organ<br />

toxicity of solubilized cancer drugs administered <strong>by</strong> intraperitoneal injection. Methotrexate and<br />

6-mercaptopurine, both FDA approved anti-cancer drugs, are known hepatotoxins. The solubility<br />

of these molecules can be increased <strong>by</strong> mixing them with a dendrimer based on melamine. C3H<br />

mice were administered subchronic doses of methotrexate or 6-mercaptopurine with and without a<br />

solubilizing dendrimer. Forty-eight hours after dosing, the mice were sacrificed, and serum was<br />

collected for biochemical analyses. The levels of alanine transaminase, ALT, were used to probe<br />

liver damage. When the drugs are encapsulated <strong>by</strong> the dendrimer, a significant reduction in hepatotoxicity<br />

is observed; ALT levels from the rescued groups (drug C dendrimer) were 27%<br />

(methotrexate) and 36% (6-mercaptopurine) lower than those of animals treated with the<br />

drug alone.<br />

Initial studies of a generation 3, cationic dendrimer bearing 24 primary amino groups revealed<br />

that this dendrimer was toxic both in vitro and in vivo. Extensive liver damage was observed when<br />

mice were challenged with subchronic doses of this dendrimer. Other groups hypothesized that<br />

surface group charge was the determining factor in predicting a dendrimer’s toxicity as cationic<br />

surface groups will electrostatically adhere to cell surface, inducing bulk adhesion that leads to the<br />

cell’s demise. Surface group modification of a generation 3, cationic dendrimer bearing 48 primary<br />

amino groups to possess anionic or neutral surface groups had a significant impact on both in vitro and<br />

in vivo toxicity. The introduction of neutral PEG groups afforded the most protection in vitro and<br />

in vivo. Shielding of the dendrimer’s charged groups attenuates the bulk adhesion of these molecules,<br />

leading to a decrease in the toxicity of these dendrimers. It appeared that PEGylation afforded the<br />

most protection against toxicity of dendrimers based on melamine. As a result, the candidate vehicle<br />

was subjected to a series of analyses. Albumin interaction studies showed that this dendrimer<br />

interacted with albumin that could have a profound effect on the dendrimer’s pharmacokinetic<br />

parameters. In vitro drug release profiles of the anti-cancer drugs methotrexate and paclitaxel<br />

revealed that methotrexate was rapidly released while paclitaxel showed a more controlled, sustained<br />

release. This discrepancy in drug release was attributed to disparities in the drug’s hydrophobicity;<br />

the more hydrophobic drug, paclitaxel, formed a stronger association with the dendrimer’s hydrophobic<br />

core. As a result, release was more controlled. In terms of the in vitro cytotoxicity of these<br />

dendrimer/drug formulations, free drug was more toxic than the dendrimer-associated drug. This was<br />

attributed to differences in the cellular uptake offree versus dendrimer-associated drug. To assess if a<br />

difference in cellular uptake existed between free and dendrimer associated substrates, cells exposed<br />

to free rhodamine B showed higher intracellular levels of dye than those cells exposed to dendrimerassociated<br />

rhodamine B. The preliminary in vivo biodistribution screen of the Cy5.5-labeled<br />

dendrimer showed that the dendrimer was still present in the body 24 h post injection while exhibiting<br />

a high degree of liver accumulation after such time (https://txspace.tamu.edu/dev-xml/bitstream/<br />

1969.1/5330/1/etd-tamu-2005A-TOXI-Neerman.pdf).<br />

Chen et al. (2004) designed a small library of dendrimers prepared from a common precursor<br />

that is available in 5 g scale in five linear steps at 56% overall yield. The precursor is a generation<br />

three dendrimer that displays 48 peripheral sites <strong>by</strong> incorporating AB4 surface groups. Manipulation<br />

of these sites provided six dendrimers that vary in the chemistry of the surface group (amine,<br />

guanidine, carboxylate, sulfonate, phosphonate, and PEGylated) that were evaluated for hemolytic<br />

potential and cytotoxicity. Cationic dendrimers were found to be more cytotoxic and hemolytic<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 543<br />

than anionic or PEGylated dendrimers. The PEGylated dendrimer was evaluated for acute toxicity<br />

in vivo. No toxicity, neither mortality nor abnormal blood chemistry, based on blood urea nitrogen<br />

levels or alanine transaminase activity was observed in doses up to 2.56 g/kg i.p. and<br />

1.28 g/kg intavenously.<br />

27.6.7 EFFECTIVE PHOTODYNAMIC THERAPY<br />

Photosensitizers play a crucial role in the photodynamic therapy (PDT) of cancer. Zhang et al.<br />

(2003) observed that the use of PIC micelles as a delivery system reduced the dark toxicity of the<br />

cationic dendrimer porphyrin, probably because of the biocompatible PEG shell of the micelles.<br />

They compared relatively low cellular uptake of dendrimer porphyrin [NH2CH2CH2NHCO]32-<br />

DPZn incorporated in the PIC micelle as was observed, yet the latter exhibited enhanced<br />

photodynamic efficacy on the Lewis Lung Carcinoma (<strong>LLC</strong>) cell line. In this study, a third-generation<br />

aryl ether dendrimer porphyrin with 32 primary amine groups on the periphery,<br />

[NH2CH2CH2NHCO]32DPZn, and pH-sensitive, polyion complex micelles (PIC) composed of<br />

the porphyrin dendrimer and PEG-b-poly(aspartic acid) were evaluated as new photosensitizers<br />

(PSs) for PDT in the <strong>LLC</strong> cell line. The preliminary photophysical characteristics of [NH2CH2-<br />

CH2NHCO]32DPZn and the corresponding micelles were investigated. Electrostatic assembly<br />

resulted in a red-shift of the Soret peak of the porphyrin core and the enhanced fluorescence.<br />

Jang et al. (2005) described a supramolecular nanocarrier of anionic dendrimer porphyrins<br />

with cationic block copolymers modified with PEG to enhance intracellular photodynamic efficacy.<br />

27.6.8 DENDRITIC GELS<br />

Luman, Smeds, and Grinstaff (2003) developed a process of high-yield convergent synthesis of<br />

dendrons, dendrimers, and dendritic-linear hybrid macromolecules composed of succinic acid,<br />

glycerol, and PEG. This convergent synthesis relies on two orthogonal protecting groups,<br />

namely, the benzylidene acetal (bzld) for the protection of the 1,3-hydroxyls of glycerol and the<br />

tert-butyldiphenylsialyl (TBDPS) ester for protection of the carboxylic acid of succinic acid. These<br />

novel polyester dendritic macromolecules are entirely composed of building blocks known to be<br />

biocompatible or degradable in vivo to give natural metabolites. Derivatization of the dendritic<br />

periphery with a methacrylate affords a polymer that can be subsequently photo-cross-linked. The<br />

three-dimensional cross-linked gels formed <strong>by</strong> ultraviolet irradiation are optically transparent with<br />

mechanical properties dependent on the initial cross-linkable dendritic macromolecule.<br />

27.6.9 LIGAND-BASED DRUG DELIVERY<br />

Architectural features of synthetic ligands were systematically varied to optimize inhibition of mast cell<br />

degranulation initiated <strong>by</strong> multivalent crossing of IgE-receptor complexes Baird et al. (2003). A series<br />

of ligands were generated <strong>by</strong> end-capping PEG polymers and amine-based dendrimers with the hapten<br />

2,4-dinitrophenyl (DNP). These were used to explore the influence of polymeric backbone length,<br />

valency, and hapten presentation on binding to anti-DNP IgE and inhibition of stimulated activation of<br />

RBL cells. Monovalent MPEG(5000)-DNP (IC(50)Z50 nM), bivalent DNP-PEG(3350)-DNP<br />

(IC(50)Z8 nM), bismonovalent MPEG(5000)-DNP(2) (IC(50)Z20 nM), bisbivalent DNP(2)-<br />

PEG(3350)-DNP(2) (IC(50)Z3 nM), and DNP(4)-dendrimer ligands (IC(50)Z50 nM) all effectively<br />

inhibit cellular activation caused <strong>by</strong> multivalent antigen, DNP-bovine serum albumin. For<br />

different DNP ligands, evidence is available for more effective inhibition because of preferential<br />

formation of intra-IgE cross-links <strong>by</strong> bivalent ligands of sufficient length, self-association of monovalent<br />

ligands with longer tails, and higher probability of binding for bisvalent ligands. They also<br />

showed that larger DNP(16)-dendrimers of higher valency trigger degranulation <strong>by</strong> cross-linking<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


544<br />

IgE-receptor complexes, whereas smaller DNP-dendrimers are inhibitory. Therefore, features of<br />

synthetic ligands can be manipulated to control receptor occupation, aggregation, and inhibition of<br />

the cellular response.<br />

27.6.10 INCREASINGLY EFFECTIVE BORON NEUTRON CAPTURE THERAPY<br />

A recent methodology in cancer treatment is the boron neutron-capture therapy (BNCT)<br />

(Hawthorne 1993). In this therapy, the generation of cytotoxic and energetic products from<br />

nuclear fission reactions of low-energy neutrons and 10 B nuclei is used to destroy malignant<br />

cells. An efficient agent for the therapy is water-soluble and has a high local density of boron<br />

clusters, requirements that are met for several synthesized boron-containing and water-soluble<br />

dendrimers. Qualmann et al. (1996a, 1996b) have, in addition, introduced antigen selectivity <strong>by</strong><br />

coupling a lysine-based boronated dendrimer to antibody fragments (Fab¢). In these agents, the<br />

attachment of a poly(ethylene glycol)(PEG) moiety is necessary to keep the conjugates watersoluble.<br />

The covalent nature of the boronated Fab¢ fragments leads to a better stability of these<br />

conjugates as compared to, for example, borate-coated polystyrene beads.<br />

Successful treatment of cancer <strong>by</strong> boron neutron capture therapy (BNCT) requires the selective<br />

delivery of 10 B to constituent cells within a tumor. The expression of the folate receptor is amplified<br />

in a variety of human tumors and potentially might serve as a molecular target for BNCT. Shukla et<br />

al. (2003) investigated the possibility of targeting the folate receptor on cancer cells using folic acid<br />

conjugates of boronated poly(ethylene glycol) (PEG) containing third generation PAMAM dendrimers<br />

to obtain 10 B concentrations necessary for BNCT <strong>by</strong> reducing the uptake of these conjugates<br />

<strong>by</strong> the reticuloendothelial system. First, they covalently attached 12–15 decaborate clusters to third<br />

generation PAMAM dendrimers. Varying quantities of PEG units with varying chain lengths were<br />

then linked to these boronated dendrimers to reduce hepatic uptake. Among all prepared combinations,<br />

boronated dendrimers with 1–1.5 PEG(2000) units exhibited the lowest hepatic uptake in<br />

C57BL/6 mice (7.2–7.7% injected dose (ID)/g liver). Therefore, two folate receptor-targeted boronated<br />

third generation PAMAM dendrimers were prepared: one containing approximately 15<br />

decaborate clusters and approximately 1 PEG(2000) unit with folic acid attached to the distal<br />

end, and the other containing approximately 13 decaborate clusters with, approximately one<br />

PEG(800) unit with folic acid attached to the distal end. In vitro studies using folate receptor<br />

(C) KB cells demonstrated receptor-dependent uptake of the latter conjugate. Biodistribution<br />

studies with this conjugate in C57BL/6 mice bearing folate receptor (C) murine 24JK-FBP<br />

sarcomas resulted in selective tumor uptake (6.0% ID/g tumor), but also high hepatic (38.8%<br />

ID/g) and renal (62.8% ID/g) uptake, indicating that attachment of a second PEG unit and/or<br />

folic acid may adversely affect the pharmacodynamics of this conjugate.<br />

27.6.11 IMPROVED MRI CONTRAST AGENTS<br />

Nanotechnology for Cancer Therapy<br />

Macromolecules conjugated with polyethylene glycol (PEG) acquire more hydrophilicity, resulting<br />

in a longer half-life in circulation and lower immunogenicity. Kobayashi et al. (2001) synthesized<br />

two novel conjugates for MRI contrast agents from a generation-4 polyamidoamine dendrimer<br />

(G4D), 2-(p-isothiocyanatobenzyl)-6-methyl-diethylenetriaminepentaacetic acid (1B4M), and one<br />

or two PEG molecules with a MW of 20,000 Da (PEG(2)-G4D-(1B4M-Gd)(62) (MW: 96 kD),<br />

PEG(1)-G4D-(1B4M-Gd)(63) (MW: 77 kD). Their pharmacokinetics, excretion, and properties as<br />

vascular MRI contrast agents were evaluated and compared with those of G4D-(1B4M-Gd)(64)<br />

(MW: 57 kD). PEG(2)-G4D-(1B4M-Gd)(62) remained in the blood significantly longer and accumulated<br />

significantly less in the liver and kidney than the other two preparations (P!0.01).<br />

Although the blood clearance was slower, PEG(2)-G4D-(1B4M-Gd)(62) was excreted more<br />

readily without renal retention than the other two preparations. The positive effects of PEG<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 545<br />

conjugation on a macromolecular MRI contrast agent were found to be prolonging retention in the<br />

circulation, increased excretion, and decreased accumulation in the organs.<br />

27.7 CONCLUSIONS<br />

Nanotechnology refers to the interactions of cellular and molecular components and engineered<br />

materials—typically clusters of atoms, molecules, and molecular fragments—at the most elemental<br />

level of biology. Such nanoscale objects—typically, though not exclusively, with dimensions<br />

smaller than 100 nm—can be useful <strong>by</strong> themselves or as part of larger devices, containing multiple<br />

nanoscale objects. At the nanoscale, the physical, chemical, and biological properties of materials<br />

fundamentally differ, and often nanotechnology can change the very foundations of cancer diagnosis,<br />

treatment, and prevention. The novel nanodevices are capable of many clinically important<br />

functions, including detecting cancer at its earliest stages, pinpointing its location within the body,<br />

delivering anti-cancer drugs specifically to malignant cells, and in killing malignant cells. As these<br />

nanodevices are evaluated in clinical trials, researchers envision that nanotechnology will serve as<br />

multifunctional tools that will not only be used with any number of diagnostic and therapeutic<br />

agents, but will also change the very foundations of cancer diagnosis, treatment, and prevention<br />

(NCI 2004).<br />

Such nanoparticulate carriers also include PEG conjugated dendrimers. These are globular,<br />

hyperbranched polymers possessing a high concentration of surface functional groups and internal<br />

cavities. These unique features make them very useful in many biomedical applications, especially<br />

as carrier molecules. There are many such technological opportunities associated with the dendrimers.<br />

Starpharma-like companies used these dendrimers to build large active compounds that<br />

present a polyvalent array to receptors, and others platform opportunities for the delivery of<br />

small molecules as a toolbox for rational drug design and other nanotechnology applications in<br />

life sciences. There are many product opportunities in dendrimers and their use such as with<br />

VivaGele, a topical microbicide gel for prevention of HIV and other STDs in women. They can<br />

act as novel chemotherapeutic agents, angiogenesis inhibitors, and they can be used for targeting a<br />

range of tropical and exotic diseases and as bio-defense agents in addition to their use for targeting a<br />

broad range of viral, respiratory, and cancer-like diseases in the future. Development of drug<br />

delivery systems with low toxicity and low cost is the key to a nanaotechnological approach to<br />

cancer treatment, including DNA-based concepts. Targeting the drugs to specific receptors<br />

exclusively expressed <strong>by</strong> tumor cells also appears to be an approach worth exploring. This biochemical<br />

approach is expected to pay rich dividends in the chemotherapy of cancer. Cytosol<br />

survival of anti-cancer agents and folate receptor-based approaches also hold promise. Carbon<br />

nanotubes offer yet another attractive possibility.<br />

REFERENCES<br />

Abou-Jawde, R. et al., An overview of targeted treatments in cancer, Clin. Ther., 25, 2121, 2003.<br />

Baird, E. J. et al., Highly effective poly(ethylene glycol) architectures for specific inhibition of immune<br />

receptor activation, Biochemistry, 42, 12739, 2003.<br />

Barar, F. S. K., Essential of Pharmacotherapeutics, S. Chand and Company Ltd, New Delhi, 2000.<br />

Bhadra, D. et al., Pegnology: A review of PEG-ylated systems, Pharmazie, 57, 5, 2002.<br />

Bhadra, D. et al., A PEGylated dendritic nanoparticulate carrier of fluorouracil, Int. J. Pharm., 257, 111, 2003.<br />

Bosman, A. W., Janssen, H. M., and Meijer, E. W., About dendrimers: Structure, physical properties and<br />

applications, Chem. Rev., 99, 1665, 1999.<br />

Brannon-Peppas, L. and Blanchette, J. O., Nanoparticle and targeted systems for cancer therapy, Adv. Drug<br />

Deliv. Rev., 56, 1649, 2004.<br />

Brigger, I., Dubernet, C., and Couvreur, P., Nanoparticles in cancer therapy and diagnosis, Adv. Drug Deliv.<br />

Rev., 54, 631, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


546<br />

Nanotechnology for Cancer Therapy<br />

Buckmann, A. F. and Morr, M., Functionalization of polyethylene glycol and mono-methoxy-polyethylene<br />

glycol, Makromol. Chem., 182, 1379, 1981.<br />

CCO Formulary, Liposomal Doxorubicin, Oct. 2003.<br />

Chapman, T. M. et al., Hydramphiphiles: Novel, linear, dendritic block copolymeric surfactants, J. Am. Chem.<br />

Soc., 116, 11195, 1994.<br />

Chari, R. V. J., Targeted delivery of chemotherapeutics: Tumor activated prodrug therapy, Adv. Drug Deliv.<br />

Rev., 31, 89, 1998.<br />

Chen, H. T. et al., Cytotoxicity, hemolysis, and acute in vivo toxicity of dendrimers based on melamine,<br />

candidate vehicles for drug delivery, J. Am. Chem. Soc., 126, 10044, 2004.<br />

Choi, Y. H. et al., Polyethylene glycol grafted poly-L-lysine as polymeric gene carrier, J. Controlled Release,<br />

54, 39, 1998.<br />

Choi, J. S. et al., Poly(ethylene glycol)-block-poly(L-lysine) dendrimer: Novel linear polymer/dendrimer block<br />

copolymer forming a spherical water-soluble polyionic complex with DNA, Bioconjugate Chem., 10,<br />

62, 1999.<br />

Choi, J. S. et al., Synthesis of a barbell-like triblock copolymer, poly(L-lysine) dendrimer-block-Poly(ethylene<br />

glycol)-block-Poly(L-lysine) dendrimer, and its self-assembly with plasmid DNA, J. Am. Chem. Soc.,<br />

122, 474, 2000.<br />

Christensen, C. A. et al., Synthesis and electrochemistry of a tetrathiafulvalene (TTF)21-glycol dendrimer:<br />

Intradendrimer aggregation of TTF cation radicals, Chem. Commun., 509, 1998.<br />

Cui, Z., Hsu, C. H., and Mumper, R. J., Physical characterization and macrophage cell uptake of mannancoated<br />

nanoparticles, Drug Dev. Ind. Pharm., 29, 689, 2003.<br />

Dagani, R., Chemists explore potential of dendritic macromolecules as functional materials, CandEN<br />

Washington, Chem. and Eng. News, June 3, 1996.<br />

De Brabander-Van den Berg, E. M. M. and Meijer, E. W., Poly (propyleneimine) dendrimers: Large-scale<br />

synthesis <strong>by</strong> heterogeneously catalyzed hydrogenations, Angew. Chem. Int. Ed. Engl., 32, 1308, 1993.<br />

Dvorak, H. F. et al., Identification and characterization of the blood vessels of solid tumors that are leaky to<br />

circulating macromolecules, Am. J. Pathol., 133, 95, 1988.<br />

Elamanchili, P. et al., Characterization of poly(D,L lactic-co-glycolic acid) based nanoparticulate system for<br />

enhanced delivery of antigens to dendritic cells, Vaccine, 22, 2406, 2004.<br />

Elizabeth, R. et al., Biological evaluation of polyester dendrimer: Poly(ethylene oxide) “bow-tie” hybrids with<br />

tunable molecular weight and architecture, Mol. Pharmacol., 2, 129, 2005.<br />

Feng, S. S. and Chien, S., Chemotherapeutic engineering: Application and further development of chemical<br />

engineering principles for chemotherapy of cancer and other diseases, Chem. Eng. Sci., 58, 4087,<br />

2003.<br />

Frechet, J. M., Functional polymers and dendrimers: Reactivity, molecular architecture, and interfacial energy,<br />

Science, 263, 1710, 1994.<br />

Freemantle, M., Blossoming of dendrimers, Chem. & Engg. News London, Science/Technology, 77, 27, 1999.<br />

Freitas, R. A. Jr., Nanomedicine, In Basic Capabilities, Landes Bioscience, Georgetown, TX, I, (http://www.<br />

nanomedicine.com/NMI.htm), 1999.<br />

Freitas, R. A., Current status of nanomedicine and medical nanorobotics, J. Comput. Theor. Nanosci.,<br />

2, 1, 2005.<br />

Gabizon, A. and Martin, F., Polyethylene glycol-coated (PEGylated) liposomal doxorubicin. Rationale for use<br />

in solid tumours, Drugs, 54, 15, 1997.<br />

Gillies, E. R. and Frechet, J. M., Designing macromolecules for therapeutic applications: Polyester dendrimerpoly(ethylene<br />

oxide) “bow-tie” hybrids with tunable molecular weight and architecture, J. Am. Chem.<br />

Soc., 124, 14137, 2002.<br />

Gillies, E. R., Jonsson, T. B., and Frechet, J. M., Stimuli-responsive supramolecular assemblies of lineardendritic<br />

copolymers, J. Am. Chem. Soc., 126, 11936, 2004.<br />

Gitsov, I. and Fréchet, J. M. J., Solution and solid state properties of hybrid linear-dendritic block copolymers,<br />

Macromolecules, 26, 6536, 1993.<br />

Gitsov, I. and Fréchet, J. M. J., Stimuli responsive hybrid macromolecules: Novel amphiphilic star copolymers<br />

with dendritic groups at the periphery, J. Am. Chem. Soc., 118, 3785, 1996.<br />

Grossfeld, G. D., Carrol, P. R., and Lindeman, N., Thrombospondin-1 expression in patients with pathologic<br />

state T3 prostate cancer undergoing radical prostatectomy: Association with p53 alterations, tumor<br />

angiogenesis and tumor progression, Urology, 59, 97, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 547<br />

Harris, J. M. et al., Synthesis and characterisation of polyethylene glycol derivatives, J. Polymer Sci. Polymer<br />

Chem., 22, 341, 1984.<br />

Harris, J. M. and Chess, R. B., Effect of pegylation on pharmaceuticals, Nat. Rev. Drug Discov., 2, 214, 2003.<br />

Hawthorne, M. F., The role of chemistry in the development of boron neutron capture therapy of cancer,<br />

Angew. Chem. Int. Ed. Engl., 32, 950, 1993.<br />

http://nano.med.umich.edu/Dendrimers.html.<br />

http://press2.nci.nih.gov/sciencebehind/nanotech.<br />

http://www.cancer.gov/clinicaltrials/results/dose-dense0604.<br />

Ishida, O. et al., Size dependent extravagation and interstitial localization of polyethylene glycol liposomes in<br />

solid tumour bearing mice, Int. J. Pharm., 190, 49, 1999.<br />

Jang, W. D. et al., Supramolecular nanocarrier of anionic dendrimer porphyrins with cationic block copolymers<br />

modified with polyethylene glycol to enhance intracellular photodynamic efficacy, Angew.<br />

Chem. Int. Ed. Engl., 44, 419, 2005.<br />

Jansen, J. F. G. A. and Meijer, E. W., The dendritic box: Shape-selective liberation of encapsulated guests,<br />

J. Am. Chem. Soc., 117, 4417, 1995.<br />

Jansen, J. F. G. A., Brabander Vanden Berg, E. M. M., and Meijer, E. W., Science, 266, 1226, 1994.<br />

Katre, N. V., The conjugation of protein with poly ethylene glycol and other polymers altering properties of<br />

proteins to enhance their therapeutic potential, Adv. Drug Deliv. Rev., 10, 91, 1993.<br />

Kim, A. et al., Pharmacodynamics of insulin in polyethylene glycol coated liposomes, Int. J. Pharmacol., 180,<br />

75, 1999.<br />

Kim, T. I. et al., PAMAM–PEG–PAMAM: Novel triblock copolymer as a biocompatible and efficient gene<br />

delivery carrier, Biomacromolecules, 5, 2487, 2004.<br />

Kim, Y. and Zimmerman, S. C., Applications of dendrimers in bio-organic chemistry, Curr. Opin. Chem. Biol.,<br />

2, 733, 1998. http://biomednet.com/elecref/1367593100200733.<br />

Kobayashi, H. et al., Positive effects of polyethylene glycol conjugation to generation-4 polyamidoamine<br />

dendrimers as macromolecular MR contrast agents, Magn. Reson. Med., 46(4), 781, 2001.<br />

Kojima, C. et al., Synthesis of polyamidoamine dendrimers having poly(ethylene glycol) grafts and their<br />

ability to encapsulate anticancer drugs, Bioconjugate Chem., 11(6), 910, 2000.<br />

Kolhe, P. et al., Drug complexation, in vitro release and cellular entry of dendrimers and hyperbranched<br />

polymers, Int. J. Pharm., 259, 143, 2003.<br />

Kreuter, J., Nanoparticles, In Colloidal Drug Delivery Systems, Kreuter, J., Ed., Marcel Dekker, Inc.,<br />

New York, 1994.<br />

Kukowska-Latallo, J. F. et al., Intravascular and endobronchial DNA delivery to murine lung tissue using a<br />

novel, nonviral vector, Hum. Gene Ther., 11, 1385, 2000.<br />

Langer, R., Biomaterials in drug delivery and tissue engineering: One laboratory’s experience, Acc. Chem.<br />

Res., 33, 94, 2000.<br />

Landers, J. J. et al., Prevention of influenza pneumonitis <strong>by</strong> sialic acid-conjugated dendritic polymers, J. Infect.<br />

Dis., 186, 1222, 2002.<br />

Liu, M., Kono, K., and Frechet, J. M., Water-soluble dendritic unimolecular micelles: their potential as drug<br />

delivery agents, J. Control. Rel., 65, 121, 2000.<br />

Luman, N. R., Smeds, K. A., and Grinstaff, M. W., The convergent synthesis of poly(glycerol-succinic acid)<br />

dendritic macromolecules, Chemistry, 9, 5618, 2003.<br />

Luo, D., A new solution for improving gene delivery, Trends Biotechnol., 22, 101, 2004.<br />

Luo, D. and Saltzman, W. M., Synthetic DNA delivery systems, Nat. Biotechnol., 18, 33, 2000.<br />

Luo, D. et al., Poly(ethylene glycol)-conjugated PAMAM dendrimer for biocompatible, high-efficiency DNA<br />

delivery, Macromolecules, 35, 3456, 2002.<br />

Maeda, H., The enhanced permeability and retention (EPR) effect in tumor vasculature: The key role of tumorselective<br />

macromolecular drug targeting, Adv. Enzyme Regul., 41, 189, 2001.<br />

Malik, N. et al., Dendrimers: Relationship between structure and biocompatibility in vitro, and preliminary<br />

studies on the biodistribution of 125I-labelled polyamidoamine dendrimers in vivo, J. Controlled<br />

Release, 65, 133, 2000.<br />

Mannisto, M. et al., Structure-activity relationships of poly(L-lysines): Effects of pegylation and molecular<br />

shape on physicochemical and biological properties in gene delivery, J. Controlled Release, 83(1),<br />

169, 2002.<br />

Maruyama, K., In vivo targeting <strong>by</strong> liposomes, Biol. Pharm. Bull., 23, 791, 2000.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


548<br />

Nanotechnology for Cancer Therapy<br />

Massing, U. and Fuxius, S., Liposomal formulations of anticancer drugs: Selectivity and effectiveness, Drug<br />

Resist. Updat., 3, 171, 2000.<br />

McNeil, S. E., Nanotechnology for the biologist, J. Leukoc. Biol., 78, 1–10, 2005.<br />

Miller, T. M., Kwock, E. W. K., and Neenan, T. X., Macromolecules, 25, 3143–3148, 1992.<br />

Moghimi, S. M. and Szebeni, J., Stealth liposomes and long circulating nanoparticles: Critical issues in<br />

pharmacokinetics, opsonization and protein-binding properties, Prog. Lipid Res., 42, 463, 2003.<br />

NCI Cancer Bulletin, National Cancer Institute, U.S. Department of Health and Human Services, National<br />

Institutes of Health, NIH Publication, Press Office No. (301) pp. 496–6641, Sep., (http://www.cancer.<br />

gov), 2004.<br />

NCI Cancer Bulletin, National Cancer Institute, U.S. Department of Health and Human Services, National<br />

Institutes of Health, NIH Publication No. 05-5498, Feb., vol. 2, (http://www.cancer.gov), 2005.<br />

Neerman, M. F. et al., Reduction of drug toxicity using dendrimers based on melamine, Mol. Pharmacol., 1,<br />

390, 2004.<br />

Newkome, G. R., Moorefield, C. N., and Vogtle, F., Dendritic Molecules: Concepts, Syntheses, Perspectives,<br />

VCH, Weinheim, 1996.<br />

Norsten, T. and Branda, N., A starburst porphyrin polymer: A first generation dendrimer, Chem. Commun.,<br />

1257, 1998.<br />

Ooya, T., Lee, J., and Park, K., Effects of ethylene glycol-based graft, star-shaped, and dendritic polymers on<br />

solubilization and controlled release of paclitaxel, J. Controlled Release, 93, 121, 2003.<br />

Padilla, D. J. et al., Polyester dendritic systems for drug delivery applications: In vitro and in vivo evaluation,<br />

Bioconjugate Chem., 13, 453, 2002.<br />

Papahadjopoulus, D. et al., Sterically stabilized liposomes improvement in pharmacokinetics and anti tumour<br />

therapeutic efficacy, Proc. Natl Acad. Sci., 88, 11460, 1991.<br />

Peppas, N. A. et al., Polyethylene glycol containing hydrogels in drug delivery, J. Controlled Release, 62, 81,<br />

1999.<br />

Phillips, W. T. et al., Polyethylene glycol modified liposomes encapsulated hemoglobin: A long circulating red<br />

blood cell substitute, J. Pharmacol. Exp. Ther., 288, 665, 1999.<br />

Qualmann, B. et al., Synthesis of boron-rich lysine dendrimers as protein labels in electron microscopy,<br />

Angew. Chem. Int. Ed. Engl., 35, 909, 1996.<br />

Qualmann, B. et al., Electron spectroscopic imaging of antigens <strong>by</strong> reaction with boronated antibodies,<br />

J. Microsc., 183, 69, 1996.<br />

Quintana, A. et al., Design and function of a dendrimer-based therapeutic nanodevice targeted to tumor cells<br />

through the folate receptor, Pharm. Res., 19, 1310, 2002.<br />

Reuter, J. D. et al., Inhibition of viral adhesion and infection <strong>by</strong> sialic-acid-conjugated dendritic polymers,<br />

Bioconjugate Chem., 10, 271, 1999.<br />

Reynolds, J. E. F., Martindale Extra Pharmacopoeia, 31st ed., Royal Pharmacol Society, London, 1996.<br />

Sahoo, S. K. and Labhasetwar, V., Nanotech approaches to drug delivery and imaging, Drug Discov. Today,8,<br />

2003, www.drugdiscoverytoday.com.<br />

Shukla, S. et al., Synthesis and biological evaluation of folate receptor-targeted boronated PAMAM dendrimers<br />

as potential agents for neutron capture therapy, Bioconjugate Chem., 14, 158, 2003.<br />

Sledge, G. and Miller, K., Exploiting the hallmarks of cancer: The future conquest of breast cancer, Eur.<br />

J. Cancer, 39, 1668, 2003.<br />

Smith, D. K. and Diederich, F. N., Functional dendrimers: Unique biological mimics, Chem. Eur. J., 4, 1353,<br />

1998.<br />

Sparks H., How miniature radiation detectors will keep astronauts safe in deep space. http://www.space.com/<br />

businesstechnology/technology/radiation_nanobots_020717.html, 2002.<br />

Storm, G., Belliot, S. O., Daemen, T., and Lasic, D., Surface modification of nanoparticles to oppose uptake <strong>by</strong><br />

the mononuclear phagocyte system, Adv. Drug Deliv. Rev., 17, 31, 1995.<br />

Suzawa, T. et al., Enhanced tumor cell selectivity of adriamycin-monoclonal antibody conjugate via a<br />

poly(ethylene-glycol)-based cleavable linker, J. Controlled Release, 9, 229, 2002.<br />

Tomalia, D. A., Naylor, A. M., and Goddard, W. A., Starburst dendrimers: Molecular-level control of size,<br />

shape, surface chemistry, topology, and flexibility from atoms to macroscopic matter, Angew. Chem.<br />

Int. Ed. Engl., 29, 138, 1990.<br />

Toyokuni, T., Hakomori, S. I., and Singhal, A. K., Bioorg. Med. Chem., 2, 1119, 1994.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


PEGylated Dendritic Nanoparticulate Carriers of Anti-Cancer Drugs 549<br />

Veronese, F. M. et al., Preparation, physicochemical and pharmcokinetic characterisation of methoxy polyethylene<br />

glycol derivatized superoxide dismutase, J. Controlled Release, 10, 145, 1989.<br />

Vincent, L. et al., Efficacy of dendrimer-mediated angiostatin and TIMP-2 gene delivery on inhibition of tumor<br />

growth and angiogenesis: In vitro and in vivo studies, Int. J. Cancer, 10, 419, 2003.<br />

Voit, B., Turner, S. R., and Mourey, T., Hyperbranched polyesters: 1. All aromatic hyperbranched polyesters<br />

with phenol and acetate end groups; synthesis and characterization, Macromolecules, 26, 4617, 1993.<br />

Wooley, K. L., Hawker, C. J., and Frechet, J. M. J., A branched-monomer approach for the rapid synthesis of<br />

dendrimers, Angew. Chem. Int. Ed. Engl., 33, 82, 1994.<br />

Zalipsky, S., Chemistry of polyethylene glycol conjugates with biologically active molecules, Adv. Drug<br />

Deliv. Rev., 16, 157, 1995.<br />

Zeng, F. and Zimmerman, S. C., Dendrimers in supramolecular chemistry: From molecular recognition to selfassembly,<br />

Chem. Rev., 97, 1681, 1997.<br />

Zhang, G. D. et al., Polyion complex micelles entrapping cationic dendrimer porphyrin: Effective photosensitizer<br />

for photodynamic therapy of cancer, J. Controlled Release, 93, 141, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


28<br />

CONTENTS<br />

Dendrimer Nanocomposites<br />

for Cancer Therapy<br />

Lajos P. Balogh and Mohamed K. Khan<br />

28.1 Composite Nanodevices .................................................................................................... 552<br />

28.1.1 Introduction ......................................................................................................... 552<br />

28.1.2 Dendrimers .......................................................................................................... 555<br />

28.1.3 Toxicity................................................................................................................ 558<br />

28.1.4 Gold and Radioactive Gold................................................................................. 558<br />

28.1.5 Dendrimer Nanocomposites ................................................................................ 559<br />

28.1.5.1 Synthesis of Gold/Dendrimer Nanocomposites................................. 561<br />

28.1.5.2 Characterization of Dendrimer Hosts and Gold<br />

Nanocomposites Used in Biodistribution Experiments..................... 562<br />

28.1.5.3 Fabrication of Nanocomposite Devices for Biologic<br />

Experiments ........................................................................................ 562<br />

28.1.5.4 Synthesis of Tritium Labeled Dendrimers......................................... 564<br />

28.1.5.5 Folate-Targeted Devices .................................................................... 564<br />

28.1.5.6 Gold Composite Nanodevice Synthesis............................................. 564<br />

28.1.5.7 Gold in the Nanocomposite Particles After Radiation<br />

Polymerization.................................................................................... 569<br />

28.1.5.8 Visualization of Gold/PAMAM Nanocomposite Uptake in<br />

Tumor Cells ........................................................................................ 569<br />

28.2 Biology .............................................................................................................................. 571<br />

28.2.1 Targeting and Nanodevice Delivery ................................................................... 572<br />

28.2.1.1 Quantitative In Vivo Biodistribution of Dendrimer and Gold<br />

Composite Nanodevices (CND)......................................................... 573<br />

28.2.1.2 Related Techniques in the Literature................................................. 573<br />

28.2.1.3 The Instrumental Neutron Activation Analysis (INAA) Method ..... 574<br />

28.2.1.4 Determination of Gold Content Using INAA—Precision<br />

and Sensitivity .................................................................................... 575<br />

28.2.1.5 Biodistribution of Positive and Neutral Surface Generation Five<br />

Dendrimers—Non-Specific Uptake ................................................... 575<br />

28.2.1.6 Biodistribution of 5 nm Gold Nanocomposites Compared with<br />

Dendrimer Templates ......................................................................... 577<br />

28.2.1.7 Biologic Differences Between Dendrimers and Gold Composite<br />

Nanodevices ....................................................................................... 577<br />

28.2.1.8 The Importance of Size on Nanocomposite Biodistribution............. 579<br />

28.2.1.9 Biodistribution Summary ................................................................... 580<br />

28.2.2 Toxicity Studies................................................................................................... 580<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

551


552<br />

28.2.2.1 Long-Term Toxicity Studies of Tritiated<br />

Dendrimer Nanoparticles ................................................................... 580<br />

28.2.2.2 Long-Term Toxicity of {Au(0)} Gold Composite<br />

Nanodevices ....................................................................................... 582<br />

28.2.2.3 Treatment of Tumors in Mouse Model Systems ............................... 582<br />

Acknowledgments ........................................................................................................................ 585<br />

Definitions .................................................................................................................................... 585<br />

References .................................................................................................................................... 586<br />

28.1 COMPOSITE NANODEVICES<br />

28.1.1 INTRODUCTION<br />

Nanotechnology for Cancer Therapy<br />

Every year, the United States alone reports more than half a million cancer-related deaths and<br />

approximately 1.3 million new cases. 3 Cancer is customarily treated with drugs (extracts, synthetic<br />

molecules, and/or biologic materials, e.g., proteins, antibodies, etc.). In an ideal case, the intervening<br />

medication should only act at the disease site and influence only those mechanisms that are<br />

causing the illness. Most of the unwanted side-effects are due to therapeutic materials that arrive at<br />

unwelcome sites (cells, tissues, organs) and influence normal mechanisms in a wrong way.<br />

Targeted drug delivery (without side effects) has been the ultimate goal of the treatment of diseases.<br />

Living objects have developed mechanisms to interact and deal with molecules and organic or<br />

inorganic materials (particles) introduced in the blood stream; there are filter organs and defense<br />

systems to eliminate the unwanted ones and keep those that are useful.<br />

Most of the anti-cancer drugs are cytotoxic—they kill cells. To use these in a clinical setting, it<br />

is mandatory to minimize the side effects, i.e., eliminate cancer cells without compromising the<br />

normal operation of other cells. The ongoing revolution in nanoscience and nanotechnology offers<br />

novel ways to treat diseases. These man-made complex objects provide an opportunity to better<br />

understand how the body works and simultaneously offer new ways of imaging and treatment.<br />

Targeted nanodevices open up many new avenues for therapy and diagnosis of cancer.<br />

Understanding biodistribution is a key to success for any medication. Biodistribution is<br />

determined <strong>by</strong> the interactions between the properties of the medicine and the living biologic<br />

system. Therefore, to ensure that a therapeutic material has only one concerted action at the<br />

molecular level, molecules/particles/devices with identical properties are needed. Mixtures of<br />

materials will have fractions with different properties, each with an individual biodistribution (of<br />

which only an envelope is observed). Different fractions with differing properties of biodistribution<br />

can lead to different actions. In simple terms, only identical particles will have a welldefined<br />

biodistribution.<br />

Because of these reasons, such materials need to be prepared that are composed of identical<br />

molecules/objects, presenting identical properties. In other words, only identical nanodevices with<br />

well-defined properties will have a defined biodistribution, making homogeneity (or, at least,<br />

narrow distribution) a fundamental requirement for a successful and reproducible targeted delivery.<br />

Nanoscience and nanotechnology encompass a broad range of areas with different degrees of<br />

maturity, namely synthesis and controlled formation of nanostructures with a variety of specific<br />

functionalities that are enabled <strong>by</strong> a widespread availability of tools to observe, characterize, and<br />

manipulate at the nanoscale level. This is a cyclic and synergistic process; improved characterization<br />

(new methods, higher spatial resolution, and higher sensitivity) leads to a better<br />

understanding of composition–structure–property relations that motivates better nanoparticle and<br />

nanostructure fabrication and allows better control of size and placement that, in turn, stimulates<br />

improved characterization.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 553<br />

Properties of nanosized particles depend not only on composition, but also on size (diameter,<br />

surface, volume), shape, structure, architecture, flexibility, conformations, etc. Materials composed<br />

of nanoparticles are characterized not only <strong>by</strong> their individual properties, but also <strong>by</strong> the distribution<br />

of those individual properties over a very large number of particles. To employ nanodevices<br />

(nanosized multifunctional particles) in therapy and imaging, it is necessary to clearly understand<br />

the relevant properties of the materials used, how the given biologic system works, and how<br />

nanodevices interact with the biologic system.<br />

To ensure the existence of these properties, appropriate characterization methods must be used<br />

that provide information about relevant properties and their distributions, including nanoparticle<br />

composition, size, shape, surface, and interior properties. At the nanoscale (1–100 nm), the surfaceto-volume<br />

ratios are very large, and surface forces (Coulomb interactions, hydrogen bonding, van<br />

der Waals forces) greatly determine the characteristics of these devices and control the interactions<br />

of these devices with biological systems.<br />

Reproducibility of nanoparticle properties is necessary to observe a reproducible biodistribution<br />

that is key to successful therapy. (Presently, most knowledge of nanoparticles is based on<br />

demonstrations of concept.)<br />

Properties and biologic behavior of nanodevices are not yet completely understood, and<br />

detailed chemical–material science–physics–biology studies are being performed to generate<br />

more materials with appropriate properties and a better understanding in the future. Empiric<br />

studies are necessary to uncover any unexpected properties of nanodevices.<br />

Radiotherapy has been used for decades as standard cancer therapy in almost all forms of<br />

cancers with varying degrees of success. One of the difficult challenges with radiation therapy is<br />

the delivery of a lethal dose of radiation to the tumor while leaving the surrounding normal tissue<br />

unharmed. Radioisotopes are often practically insoluble in water or body fluids, and they are rarely<br />

used as elements. Typically radioisotopes are attached to an appropriate carrier molecule. Tumordirected<br />

antibody (or peptide) radiation therapy has been used to achieve this goal with limited<br />

success. In this technique, the radiation emitter is attached to the targeting antibody or peptide,<br />

often <strong>by</strong> a cage-type carrier, that is linked to the targeting molecule. The caged carrier can carry<br />

only a limited number of radioactive atoms/molecules and usually requires a linker to prevent<br />

destruction of the targeting molecule. This creates a fundamental limitation for the dose that can be<br />

delivered per targeting molecule that may limit the ability to successfully treat solid tumors with<br />

radioactive antibody approaches. For any targeted dose delivery approach, it has to be demonstrated<br />

that the treatment will be effective if the dose is directly delivered to the tumor<br />

Composites are physical mixtures or two or more components that display improved properties<br />

in one or more areas as compared to their individual bulk constituents. They are widely used in<br />

many diverse industries and form the basics of engineering plastics, structural adhesives, and<br />

matrices. Nanocomposites can be defined as having at least one component with at least one of<br />

their dimensions less than 100 nm. Therefore, nanocomposite properties will be strongly influenced<br />

<strong>by</strong> extensive interfacial interactions between the components at the interphase, and the rule of<br />

mixtures may fail to provide estimates of nanocomposite properties.<br />

Guest–host nanodevices composed of therapeutic (or imaging) materials carried in micelles,<br />

liposomes, polymeric nanocarriers, dendrimers, nanoemulsions, nanocomposites, etc., are a wellstudied<br />

class of multifunctional nanomaterials with several potential medical uses, including<br />

targeted drug-delivery, cancer imaging, and therapy. Typically, the common feature in these<br />

structures is that the carried material is not bound to the carrier with covalent bonds, but it is<br />

constrained <strong>by</strong> physical forces. Composition, 3-D structure, and surface chemistry of the carrier are<br />

critical parameters for targeted delivery. Charge and size of the nanodevices determine their in vivo<br />

biodistribution, and there<strong>by</strong>, the efficacy for imaging and therapies.<br />

Composite nanodevices possess chemical and physical (spectral) properties, both the inorganic<br />

and the organic components, and physical and biologic interactions of the nanoparticles with each<br />

other and with the environment (solubility, biocompatibility, etc.) are dominated <strong>by</strong> the size, shape,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


554<br />

Nanotechnology for Cancer Therapy<br />

and contact surface of the host. As a consequence, in hybrid composite nanodevices, a specific<br />

nanoparticle property (e.g., radioactivity belonging to the inorganic guest) can be manipulated as if<br />

it were a property of the organic host (e.g., dendrimer).<br />

Targeted composite nanodevices are constructed <strong>by</strong> covalently binding targeting moieties to<br />

the organic template that dominates the surface of the nanocomposite. The surface can be chemically<br />

modified with various substituents, including proteins to achieve the necessary organ/tissue<br />

specificity and create targeted composite nanodevices. The interaction of these nanodevices with<br />

the biologic environment (solubility, biocompatibility) is controlled <strong>by</strong> the contact surface of the<br />

host molecule, i.e., <strong>by</strong> their size and surface properties. The separation of overall properties from<br />

the surface chemistry (targeting, charge, and size) is a unique advantage in the optimization of these<br />

nanodevices because properties of the components can be individually selected and optimized. 4<br />

Nearly monodisperse gold/dendrimer nanocomposites can be synthesized in various predetermined<br />

sizes, and their interaction with biological objects may be adjusted <strong>by</strong> modifying their surface<br />

properties. They can readily penetrate cells and are easy to observe <strong>by</strong> various methods. 5 Composite<br />

nanoparticles with either cationic, anionic, neutral, lipophilic, lipophobic, or mixed surfaces can be<br />

created <strong>by</strong> encapsulating passive or active therapeutic agent(s) that adsorb or emit radiations,<br />

respectively. Guests in radioactive nanocomposite devices may be metals or other compounds<br />

containing isotopes of medical importance permitting imaging labels, (e.g., 125 I, 3 H) or radiation<br />

therapy, for example, 198 Au that can deliver b-radiation to tumors.<br />

Using targeted nanodevices allows one to deliver various functionalities via a single biohybrid<br />

nanoparticle, making these nanodevices a potentially attractive tool in radiation medicine. An<br />

analysis of tumor type should allow for the selection and rapid immobilization of the appropriate<br />

radioactive isotope in the pre-made template that can then be efficiently delivered to specifically kill<br />

tumor cells with minimal collateral damage. The charge, size (i.e., surface chemistry), composition,<br />

and 3-D structure of these nanoparticles are critical in determining their in vivo biodistribution, and<br />

therefore, the efficacy of nanodevice imaging and therapies.<br />

Some of the fundamental problems that can be overcome <strong>by</strong> the use of targeted composite<br />

nanodevices are solubility and dose delivered. First, many radioisotopes are often practically<br />

insoluble in water or body fluids, but they may become readily available in nanocomposites, i.e.,<br />

in different sizes, contents, and surfaces. Second, use of composite nanodevices permits at least a<br />

log-fold higher delivery of radioactivity than that available with current radioactive antibody<br />

therapies. As opposed to labeled antibodies, at least 1–1000 radioisotope nuclei can be encapsulated<br />

in one nanocomposite device 6. Therefore, radioactive nanocomposite particles will be able to<br />

deliver therapeutic dose amounts to tumor cells. (Recent data indicate that even one radioactive<br />

alpha-particle decay could trigger cell death if it occurs in the cell.) 7<br />

The specific advantage of the targeted dose delivery <strong>by</strong> composite nanodevices approach over<br />

intracellular drug delivery or simple diffusion of drug into cells is that cell surface delivery is<br />

sufficient and no internalization is needed. As the specific activity of one composite { 198 Au}<br />

nanodevice is extremely small (in the range of 1E-15 Ci/nanoparticles, i.e., 3.7E-25 Bq/nanoparticle),<br />

single particles delivering b-radiation at the cell surface (low-level non-specific uptake<br />

<strong>by</strong> T-cells) are harmless, but concentrations above a certain critical level (i.e., when large number<br />

of particles are delivered to specific cells or tissues) can be therapeutic.<br />

Practically, for any targeted delivery approach, it has to be demonstrated that the treatment is<br />

effective if the therapeutic material accumulates in the tumor. These conceptual experiments must<br />

be followed <strong>by</strong> biodistribution studies and the optimization of devices.<br />

In a proof-of-concept research, { 198 Au} nanodevices were tested, but conceptually, a number of<br />

other nuclides with the desired specific activity could be used (allowing the use of gamma, beta,<br />

alpha, or combinations of these radiations in the future). As gold is non-toxic and the hot and cold<br />

gold isotopes are chemically identical, biodistribution studies may conveniently be done with<br />

nanocomposites containing non-radioactive gold.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 555<br />

28.1.2 DENDRIMERS<br />

The term dendrimer may refer to a mathematical structure, a chemical structure of a single<br />

molecule of a certain class of macromolecules, or to a certain material (usually a product of a<br />

specific synthesis procedure) 8,9 (Figure 28.1). A dendrimer material is typically a mixture of very<br />

similar polymer molecules.<br />

A number of characteristic general properties of dendrimers can be concluded from the<br />

dendritic structure itself as a result of symmetry and maximal degree of branching. These characteristics<br />

are the function of the core structure, the connector length, the degree of branching, etc.,<br />

and they appear in all the families (for example, the tendency of dendrons to move in a concerted<br />

fashion). 10–13<br />

Dendrimer molecules are symmetric molecules with possible molecular masses of up to several<br />

million daltons (i.e., dendritic structures built from atoms) that contain connectors and branching<br />

units composed of repetitive shell structures around a core following a predefined molecular motif.<br />

Dendrimer molecules always have specific chemistry and specific properties.<br />

There are many ways to construct polymer molecules with dendritic architecture, depending on<br />

which part of the macromolecule is dendritic (and what different authors consider to be a<br />

dendrimer). To date, more than 200 various dendritic structures have been reported in the literature.<br />

The major synthesis strategies are the divergent route that leads to higher yields with lower<br />

dendritic purity. 14 Convergent strategies 15 usually result in more uniform molecules but with<br />

lower overall yields (a great number of other strategies have been also reported). Dendrimer<br />

molecules are well-defined and highly symmetrical molecules containing a large number of regularly<br />

spaced internal and external functional groups. (Dendritic molecules are built to maximize<br />

symmetric branching.) As a result, the interior of a dendrimer may considerably differ from its<br />

exterior and may be hydrophilic or hydrophobic, depending on the design and synthesis route.<br />

Constructing a dendritic network from atoms necessarily invokes specific properties because<br />

the atoms themselves have specific properties. Dendrimer families contain a certain group of<br />

molecular motifs (PAMAM, PPI, polylysine dendrimers, polyether dendrimers, polyacetylene<br />

dendrimers, etc.). They are synthesized on different synthetic principles and use different routes;<br />

therefore, they have different chemistries. Dendrimer families may be very different.<br />

Generations and their subclasses. Within any families there are low, high, and middle generation<br />

dendrimers. Molecules of low generation dendrimers (LGD) structurally are relatively small<br />

organic molecules with fully flexible loose networks and with a few interactive sites (e.g., the<br />

FIGURE 28.1 (a) Dendritic structure (three-functional core, bifunctional branching, persistent angles) and a<br />

computer model of an ideal PAMAM dendrimer molecule (b). (The left panel has been reproduced from<br />

Bielinska, A., Eichman, J. D., Lee, I., Baker, J. R., and Balogh, L., J. Nanopart. Res., 4, 395–403, 2002. With<br />

permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


556<br />

Nanotechnology for Cancer Therapy<br />

chemical accessibility of the functions are roughly equal). High generation dendrimers (HGD) are<br />

essentially highly charged organic nanoparticles where the accessibility of surface groups is much<br />

higher than of the interior. These organic nanoparticles have a well-defined generational size and a<br />

persistent surface, properties that may be quite different from its interior. Medium generation<br />

dendrimers (MGD) are transitional and may display properties of both groups.<br />

A dendrimer material is typically a mixture of very similar polymer molecules as only perfect<br />

reactions (complete conversion with 100% yield) or perfect purification would lead to perfect<br />

structures in the consecutive synthesis steps. Consequently, practical dendrimer materials are<br />

oligomers or macromolecules with narrow polydispersity, and they always contain molecules<br />

with imperfect structures. However, the high level of synthetic control and effective purification<br />

between the synthesis steps allows the synthesis of defined polymers with very low polydispersities.<br />

Material properties are usually different for different generations; for example, in viscosity<br />

measurements, LGD materials can be described as draining networks, whereas HGD materials<br />

behave as ideal Newtonian fluids. There may be measurable differences between batches of individually<br />

manufactured or synthesized dendrimer materials; therefore, one must be very careful to<br />

generalize experimental results to the class of dendrimers. Dendritic properties of a given material<br />

can neither be assessed nor evaluated without taking into account the family, generation, composition,<br />

various properties, and their distribution (Figure 28.2).<br />

Polyionic water-soluble dendrimers are those where tertiary nitrogen provides the branching.<br />

Poly(amidoamine) (PAMAM) dendrimers contain beta-alanine subunits and can be synthesized to<br />

be biofriendly. 16–19 They are well characterized and are also commercially available. PAMAM<br />

dendrimer molecules undergo changes in size, shape, and flexibility as a function of increasing<br />

generations (from a small branching molecule at generation 0–2 (LGD) through flexible macromolecules<br />

(MGD: G3, G4, G5–G6) to dense organic particles behaving as hard spheres at<br />

generations seven and above) (Figure 28.2). Branching structure of MGD and HGD PAMAMs<br />

can efficiently entrap therapeutic molecules. 20–22 They have been used as delivery vehicles for<br />

oligonucleotides and antisense oligonucleotides and as probes for oligonucleotide arrays. 23,24<br />

PAMAM dendrimers are not broken down <strong>by</strong> enzymes in vivo and generally are removed from<br />

the bloodstream <strong>by</strong> the filter organs. 25–27<br />

Diversities that exist in dendrimer materials are of generational, skeletal, and/or substitutional<br />

origin. It is common that traces of various generations are present in materials (Figure 28.3a, peaks<br />

A–B–C in SEC-MALLS); the individual molecules may or may not be perfectly dendritic (see:<br />

skeletal diversity), and less than complete substitution of the end groups results in slightly different<br />

substitutions on the individual molecules (because of the random nature of chemical reactions, i.e.,<br />

FIGURE 28.2 Size and shape of ethylenediamine core PAMAM dendrimers as a function of generation (fully<br />

protonated structures in water, i.e., maximally extended conformations). Dendritic polymers provide a multifunctional<br />

surface with a variable size and tunable compatibility. (Reproduced from Bielinska, A., Eichman, J.<br />

D., Lee, I., Baker, J. R., and Balogh, L., J. Nanopart. Res., 4, 395–403, 2002. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 557<br />

Relative scale<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

B<br />

C A<br />

LS<br />

LV<br />

RI<br />

–0.2<br />

30 31 32 33 34 35 36 37 38<br />

(a)<br />

Time (min)<br />

substitutional diversity). 28 The most important properties of dendrimers are mass and charge<br />

(Figure 28.3). In case of HGD (e.g., GO6 in the case of PAMAMs) that are dense-packed,<br />

contribution to the surface may be solely due to terminal groups. In the more open structures of<br />

lower generations, other functional groups also contribute in an increasing fashion as the generational<br />

number decreases.<br />

Dendrimers have a very rich synthetic chemistry describing how molecules with dendritic<br />

architectures can be synthesized and/or modified (to date, more than 5000 papers have been<br />

published on dendrimers). 8,9 Large dendritic structures may be synthesized <strong>by</strong> reacting dendrimers<br />

as building blocks (e.g., as core and shell reagents) with each other (Figure 28.4). These molecules<br />

have been termed tecto-dendrimers. 21,29,30<br />

Among other means, PAMAM dendrimers have been proposed as versatile vehicles of drug<br />

transport to specific organs and tissues because of their unique external and internal functionalities<br />

that can be altered according to need. These spherical macromolecules have modifiable surface<br />

functionalities offering a versatile mode to covalently attach drugs, diagnostic/imaging modules,<br />

and/or targeting moieties, and they allow the control of effective surface charge. However, the<br />

expensive starting materials, the required multistep synthesis, and subsequent characterization of<br />

the multifunctional nanodevices often present a serious challenge. 31 This complexity may be<br />

Relative scale<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

10.54 min<br />

–10<br />

0 5 10 15 20<br />

(b)<br />

Time/min<br />

E5.NH2<br />

25 30 35<br />

FIGURE 28.3 (a) Size exclusion chromatogram (SEC-MALLS) showing mass (MnZ28,940) and polydispersity<br />

(PDIZ1.067) A: G4 trailing generation; B: G5 main component; C: dimers of G5 present in the<br />

material; and (b) a capillary electropherogram (CE) illustrating the almost identical charge/mass ratio for an<br />

actual ethylenediamine core primary amine-terminated generation five PAMAM dendrimer material.<br />

G y<br />

Core reagent<br />

(NH 2 )<br />

−<br />

G x<br />

Shell reagent<br />

(excess)<br />

(1) Self assembly<br />

(Equilibration)<br />

CO2H (2) Covalent bond<br />

formation<br />

G x<br />

G x<br />

G x<br />

G y<br />

G x<br />

G x<br />

G x<br />

Generation 1 core-shell<br />

Tecto(dendrimer)<br />

Gz<br />

(NH 2 )<br />

G z<br />

G z<br />

G z<br />

G z<br />

G x<br />

G x<br />

G z<br />

Gz<br />

G x<br />

G y<br />

G x<br />

G z<br />

G z<br />

G x<br />

G x<br />

G z<br />

G z<br />

G z<br />

G z<br />

Generation 1 core-shell<br />

Tecto(dendrimer)<br />

FIGURE 28.4 Scheme of tecto-(dendrimer) synthesis. The method is used to make large templates and supramolecular<br />

nanoparticles for multi-functional composite nanodevice fabrication.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


558<br />

considerably simplified using hybrid composite nanoparticles templated <strong>by</strong> monodisperse dendrimers<br />

(Figure 28.5).<br />

Polyionic dendrimers are close to being optimal templates for these composite nanoparticles<br />

because the better the templates, the more homogenous the resulting composite particles become.<br />

PAMAM and PPI can interact with a wide variety of ions and in situ synthesized compounds. 32–38<br />

These tertiary nitrogens act as focal points for pre-orientation either of cations to coordinatively<br />

bind through donor–acceptor interactions of the lone electron pair or of anions that can electrostatically<br />

bind to the positive charge of the protonated –NZligands. Multiple binding within the<br />

large molecule (a nanoscopic size pseudo-phase) also contributes to the unusually strong binding<br />

and high binding capacity. 39,40 Therefore, using PAMAM dendrimers as templates, the pre-orientation<br />

and dispersion of the inorganic components are controlled <strong>by</strong> the appropriate dendritic<br />

polymer host 32,34,35 and the size, size-distribution, shape, and surface of the resulting composite<br />

particle can be designed. Countless combinations are possible between metal cations or anions<br />

and dendrimers, but examples will be limited to aqueous solutions of generation (g4–g6) ethylenediamine<br />

core poly(amidoamine) (PAMAM) dendrimers and gold in the form of complex<br />

[AuCl 4] K (tetrachloroaurate) anions. 41,42<br />

28.1.3 TOXICITY<br />

+MY<br />

+X<br />

D [(MY) n –D] {(MX) n –D}<br />

–Y<br />

Dendrimer<br />

Dendrimer biocompatibility and toxicity have been recently reviewed, and details can be found in<br />

the literature. 16 Briefly,ininvitroexperiments, it was observed 43 that amine-terminated G5<br />

PAMAM dendrimers decreased the integrity of the cell membrane and allowed the diffusion of<br />

cytosolic proteins out of the cell and dye molecules into the cell. Although neither G5 amine- nor<br />

acetamide-terminated PAMAM dendrimers were cytotoxic up to 500 nM concentration, the dose<br />

dependent release of the cytoplasmic proteins lactate dehydrogenase (LDH) and luciferase (Luc)<br />

indicated the presence of holes. The induction of permeability caused <strong>by</strong> the amine-terminated<br />

dendrimer was not permanent, and leaking of cytosolic enzymes returned to normal levels upon<br />

removal of the dendrimers. Diffusion of dendrimers through holes is sufficient to explain the nonspecific<br />

uptake of positively charged PAMAM dendrimers into cells and is consistent with the lack<br />

of uptake of neutral surface PAMAM dendrimers.<br />

28.1.4 GOLD AND RADIOACTIVE GOLD<br />

Nanotechnology for Cancer Therapy<br />

Complex/Salt Complex/Sal t<br />

Nanocomposite<br />

FIGURE 28.5 General scheme of dendrimer nanocomposite synthesis through complex formation followed<br />

<strong>by</strong> reactive encapsulation. (Reproduced from Bielinska, A., Eichman, J. D., Lee, I., Baker, J. R., and Balogh,<br />

L., J. Nanopart. Res., 4, 395–403, 2002. With permission.)<br />

Noble metals, especially gold, have many applications in optical and biological sciences as a result<br />

of their stability. Metallic gold is usually produced <strong>by</strong> reduction using Na-citrate, citric acid,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 559<br />

hydrazine, hydrogen, or electrochemical deposition, and the color of these particles is the function<br />

of size and shape of the metal domains formed. 44<br />

Gold-based therapeutic agents have been recently reviewed. 45 Gold nanoparticles 46 and nanoshells<br />

47–49 have recently been successfully tested for photothermal therapy of cancer. Radioactive<br />

gold has been approved <strong>by</strong> FDA for radiation treatment and the use of 198 Au colloids for cancer<br />

therapy has already been attempted in the 1950s. 45,50–52 Radioactive 198 Au is used in interstitial<br />

brachytherapy for treatment of prostate cancer, initially in the form of gold colloids (micron-sized<br />

particles), 53 and later with seed technique. 54<br />

With biopsy results approaching a negative rate of 80%, and at five years, a cancer specific<br />

survival of 100% for Stages A and B1, 90% for Stage B2, and 76% for Stage C, this form of<br />

treatment offers an effective and well-tolerated alternative mode of therapy for patients with<br />

localized prostate cancer. Present evaluation of the accumulated Au(0) related biodistribution<br />

data is almost impossible because of the partial or full lack of reliable particle characterization<br />

describing size and/or surface properties 51,52 and literature thereof.<br />

Use of radioactive gold cluster immunoconjugates (containing 11–33 Au atoms) was reported<br />

in the early 1990s, but because of the cumbersome preparation and insufficient delivered dose, the<br />

research was discontinued. 55 Recently, 30–34 nm multifunctional, TNF-conjugated gold nanoparticles<br />

were prepared and tested for targeting solid tumors in mice. 36,41,56,57<br />

28.1.5 DENDRIMER NANOCOMPOSITES<br />

Dendrimer templated nanocomposites are synthesized <strong>by</strong> reactive encapsulation. 58,59 According to<br />

the general concept of DNCs, 32 precursors to a desired product are pre-organized <strong>by</strong> an appropriately<br />

selected dendrimer in the first step. This pre-organization is followed <strong>by</strong> in situ chemical<br />

reaction(s) or physical treatment (irradiation, etc.) that generates reaction products immobilized in<br />

the nanoscopic size dendrimer molecules (Figure 28.5). This procedure yields dispersed small<br />

domains of guest molecules that are integrated with the dendrimer molecule(s) without creating<br />

covalent bonds between the dendrimer and the topologically entrapped matter. Dendrimer<br />

templated nanocomposites may be constructed with flexible architectures and in<br />

variable compositions.<br />

Structure of nanocomposite particles was found to be the function of the dendrimer structure<br />

and surface groups as well as the formation mechanism and the involved chemistry. Based on the<br />

locality of the guest atoms, three different types of single nanocomposite architectures have been<br />

identified such as internal (I), external (E), and mixed (M) type nanocomposites where the inorganic<br />

matter either entrapped in the interior, in the exterior, or in both part of the dendrimer, respectively.<br />

Dominant properties of the composite nanoparticles approach the dendrimer properties with<br />

decreasing M/D ratios (i.e., with the molar ratio of the inorganic reactants and the organic<br />

template). Stable internal structures form when medium and high generation polyionic dendrimers<br />

with a weak binding periphery are employed with proper chemistry and at relatively low metal/dendrimer<br />

ratios. (In this case, the inorganic domains are located in the interior of the dendrimer<br />

templates.) Single composite nanoparticles may be used as building blocks in the synthesis of<br />

nanostructured materials and devices. Details of experimental procedures and analytical techniques<br />

can be found in previous studies. 41<br />

Dendrimer nanocomposites can be constructed in precise sizes and well-defined surfaces with<br />

various covalently attached targeting moieties and with variable content. DNCs display stability<br />

and robustness in addition to properties represented <strong>by</strong> their individual components as a result of the<br />

molecular level’s mixing of uniform macromolecules and nanoparticles. The unique advantage of<br />

this approach is that host and guest components can be separately selected, developed, and<br />

optimized. As a result of various synthetic options, the interior and/or the exterior of the host<br />

can be cationic, anionic, or non-ionic, depending on their termini and interior functionalities and<br />

the pH.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


560<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 28.6 Scheme of assembling templates and synthesis of nanocomposites of well-defined sizes and<br />

surfaces (a) and TEM image of an {Au} nanocomposite single particle and a dimer of {Au-PAMAM} n multiparticle<br />

(b). (Reproduced from Bielinska, A., Eichman, J. D., Lee, I., Baker, J. R., and Balogh, L., J. Nanopart.<br />

Res., 4, 395–403, 2002. With permission.)<br />

Preparation of modular composite structures is possible either (a) <strong>by</strong> the synthesis of modular<br />

templates from simple dendrimer molecules (i.e., using tecto-dendrimers), or alternatively, (b)<br />

synthesizing of modular nanocomposites from single nanocomposite units. Variation of nanocomposite<br />

size is possible <strong>by</strong> using larger templates such as tecto-dendrimers or <strong>by</strong> combining existing<br />

nanocomposite particles. Alternatively, the size of nanoparticles may also be varied <strong>by</strong> reacting the<br />

surface of premade nanoparticles (Figure 28.6a and Figure 28.6b). Stable internal composite<br />

structures cannot be formed using linear macromolecules, only from dendrimers. This is the substructure<br />

that has great potential in combining the properties of organic and inorganic materials for<br />

biomedical applications. For these nanocomposite structures, the particle surface is supplied <strong>by</strong> the<br />

macromolecule; therefore, the solubility and surface compatibility of the nanocomposite particles is<br />

very similar to the template dendrimer macromolecules (Figure 28.7). These DNC structures are<br />

most frequently amorphous or nearly amorphous. 60<br />

Radioisotopes are often practically insoluble in water or body fluids; however, the nanocomposites<br />

are soluble in different sizes, contents, and surfaces. An analysis of tumor type should allow<br />

for the selection of a targeter, allowing rapid immobilization of the appropriate radioactive isotope<br />

in the pre-made template. Targeted composite nanoparticles can then be efficiently delivered to<br />

specifically kill tumor cells with minimal collateral damage.<br />

As a basis for a future targeted 198 Au-based image-guided therapy, the effects of size and<br />

surface charge on the in vivo biodistribution, excretion, and toxicity of gold/dendrimer<br />

FIGURE 28.7 (a) Computer simulation of a dendrimer nanocomposite; (b) HRTEM of a gold nanocomposite<br />

particle containing 14 Au atoms. (Reproduced from Khan, M. K., Nigavekar, S. S., Minc, L. D., Kariapper, M.<br />

S., Nair, B. M., Lesniak, W. G., and Balogh, L. P., Technol. Cancer Res. Treat. 4(6), 603–613, 2005. With<br />

permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 561<br />

nanocomposites in a mouse melanoma tumor model system were studied. The nanodevices tested<br />

include 5 and 22 nm primary amine surface, 5 nm carboxylate surface, and 11 nm carboxylate<br />

surface gold/dendrimer nanocomposites that display positive and negative surface charges at biological<br />

pH values. Gold nanocomposites offer the high electron density and radioactivity of the<br />

immobilized gold atoms, and the interactions of the composite particle with the surrounding<br />

biologic environment are determined <strong>by</strong> the surface. Biologic fate can be influenced <strong>by</strong> increasing<br />

or decreasing the number and/or strength of interacting forces on the nanparticle surface <strong>by</strong> changing<br />

the number and character of the dendrimer terminal groups.<br />

28.1.5.1 Synthesis of Gold/Dendrimer Nanocomposites<br />

To synthesize gold/dendrimer composite nanodevices in nearly uniform sizes and with close to<br />

identical properties, polyionic PAMAM dendrimer templates with low (4–10) M/D ratios were<br />

used. This method leads to well-controlled and uniform composite particles. The product is a<br />

synthetic dendrimer-dominated material that is water-soluble and stable. These hybrid nanoparticles<br />

have tunable properties; they can be made to be non-immunogenic, and they are similar in size<br />

to fundamental proteins present in the blood. 61<br />

Synthesis of {Au(0)} gold/dendrimer composite nanoparticles from dendrimer templates<br />

(Figure 28.5) involves two synthesis steps: (a) forming a PAMAM-tetrachloroaurate salt, and (b)<br />

the transformation of [AuCl4]K through Au 3C ions to Au(0). Hydrolysis of [AuCl4] K is a complex<br />

process 62 and easily occurs in aqueous solutions of PAMAMs as the nitrogens neutralize HCl, the<br />

hydrolysis <strong>by</strong>product. Reduction is usually performed either <strong>by</strong> adding external reducing agents<br />

(N2H4 or NaBH4) 41,57 or utilizing the reductive nature of terminal groups of PAMAMs. 63<br />

In the two extreme cases, (depending on the reaction conditions) the dendrimers may either<br />

encapsulate the colloidal gold, 57 or alternatively, gold nanocrystals may grow in the interior of the<br />

PAMAM molecules. 6 Reduction of Au 3C ions <strong>by</strong> the PAMAM itself may be accelerated <strong>by</strong><br />

applying UV-irradiation. 64 The usual size of dendrimer templated nanocomposite particles is<br />

dZ3–10 nm, depending on the size of the template, the gold/dendrimer ratio, and the type of<br />

the composite particle. 41,65 Smaller (LGD) PAMAMs provide larger, multiparticle composites<br />

because of their open structures, whereas for MGD and HGD (G5–G9) PAMAMs and at low<br />

metal/dendrimer ratios, the composite particles preserve the size of the template. 66,67<br />

Controlled fabrication of large nanocomposite particles requires large templates. Synthesis of<br />

positively charged large (dO10 nm) PAMAM dendrimer structures, (e.g., generation 9 PAMAM<br />

with an MnZ467,000 and dZ11.4 nm) the routinely used divergent route requires several<br />

months. The multi-step synthesis and purification (altogether 18 step for G9) results in expensive<br />

(presently $16,500/g), materials. 68 Applying tecto-dendrimers 69–71 as templates is a more favorable<br />

approach, but tecto-PAMAMs can provide only methylester or carboxylate terminated (i.e., negatively<br />

charged) composite nanoparticles.<br />

It has been found that neutron/gamma irradiation also induces elemental gold formation in<br />

aqueous solutions or gels (submitted for publication). A novel and simple procedure to fabricate<br />

{( 198 Au(0)n-PAMAM}C positively charged radioactive gold nanocomposites in dZ10–30 nm<br />

sizes directly from PAMAM tetrachloroaurate salts or dZ5 nm {Au(0) n-PAMAM_E5.NH 2} kC<br />

nanocomposites <strong>by</strong> simultaneously utilizing both the neutron and gamma-radiation have been<br />

devised and developed. This method was used to fabricate positively charged radioactive gold<br />

nanocomposites in sizes between 10 nm and 30 nm, also to be used as dose delivery agents (Nano<br />

Letters, <strong>2006</strong>, submitted).<br />

Particle size, mass, and charge of dendrimer nanocomposites may be very similar to their<br />

templates. Shown in Figure 28.8 is the PAGE electropherogram of different generations of<br />

amine-terminated dendrimer templates and the corresponding dendrimer nanocomposites. These<br />

Au-nanocomposites were templated <strong>by</strong> polycationic (amine terminated) PAMAM dendrimers of<br />

different generations with the same molar ratio of nitrogen ligand/gold atom and were characterized<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


562<br />

FIGURE 28.8 PAGE electropherograms of several polycationic PAMAM dendrimer templates and their<br />

corresponding gold-dendrimer nanocomposites. Observe the similarity between the movement patterns of<br />

the dendrimer templates and the nanocomposites, indicating similar charge/mass ratios for the templates<br />

and the gold nanocomposites (D nitrogen ligand/Au ratioZ4.5–5.2). (Reproduced from Khan, M. K., Nigavekar,<br />

S. S., Minc, L. D., Kariapper, M. S., Nair, B. M., Lesniak, W. G., and Balogh, L. P., Technol. Cancer<br />

Res. Treat. 4(6), 603–613, 2005. With permission.)<br />

after a prolonged storage, i.e., in thermodynamic equilibrium. The amine-terminated golddendrimer<br />

nanocomposites and their respective dendrimer templates all display very similar<br />

migration patterns. (Au-nanocomposites display somewhat lower electrophoretic mobility probably<br />

because of the slightly higher mass of the composite particles.) 65<br />

In any case, using a good quality template is essential for success. Templates can ease the<br />

difficulty of defining size and shape, but the quality of the product will be determined <strong>by</strong><br />

the template’s precision. In short, the product cannot be better than the template and the procedure<br />

used. In an ideal case, one would have a perfect template and a subsequent perfect procedure to get<br />

a homogenous and uniformly sized nanomaterial of a predefined shape. Any imperfection in the<br />

template or the procedure results in a broadening distribution of composition and geometry (and,<br />

subsequently, of the properties) in the product.<br />

28.1.5.2 Characterization of Dendrimer Hosts and Gold Nanocomposites Used in<br />

Biodistribution Experiments<br />

Determination of fundamental properties (composition, size, and charge) is vital for reproducibility<br />

prior to initiating any biological experiment. A variety of analytical techniques to were applied to<br />

investigate the purity, properties, and structural characteristics of dendrimer hosts and host-guest<br />

nanocomposites, including polyacrylamide gel electrophoresis (PAGE), capillary electrophoresis<br />

(CE), 72 size exclusion chromatography (SEC), and HPLC, 73,74 UV–visible and fluorescence<br />

spectrophotometry, acid–base potentiometric titration, mass spectrometry (ESI-MS and MALDI-<br />

TOF techniques), and various NMR techniques. 65,75,76 (available at www.mrs.org). 77 Gold/dendrimer<br />

nanocomposites have been characterized <strong>by</strong> UV–visible and fluorescence spectrophotometry,<br />

NMR, DLS (dynamic light scattering), and TEM. Particle structure of the neutron-irradiated<br />

gold nanocomposites have also been studied (after isotope decay) <strong>by</strong> methods described above<br />

in order to evaluate the effect of direct activation on the stability of the polymer host. 4,42,61<br />

28.1.5.3 Fabrication of Nanocomposite Devices for Biologic Experiments<br />

Nanotechnology for Cancer Therapy<br />

In these studies, non-targeted and folic acid targeted hosts were synthesized and characterized (i.e.,<br />

dZ5 nm) 3 H labeled PAMAM dendrimers with positive and neutral surface charges in normal and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 563<br />

TABLE 28.1<br />

Molecular Characteristics of Polyanionic PAMAM Dendrimer Templates<br />

Mw a<br />

Mn b<br />

Mw b<br />

Polydispersity b<br />

PAMAM E3.SAH E5.SAH E5(E3.SAH) 9<br />

10,109 41,626 147,542<br />

6300 38,400 152,200<br />

9790 38960 182,100<br />

1.022 1.045 1.196<br />

Theoretical no. of COOH groups a<br />

32 128 256<br />

Practical no. of carboxylate groups c<br />

30.6 113.2 296<br />

Charge/mass (mol/g) a,d<br />

K3.18!10 K3<br />

K3.08!10 K3<br />

Charge/mass (mol/g) b,c<br />

K3.32!10 K3<br />

K2.90!10 K3<br />

a<br />

Theoretical molecular weight assuming complete conversions.<br />

b<br />

As measured <strong>by</strong> SEC at pH 2.74.<br />

c<br />

Measured <strong>by</strong> potentiometric titration.<br />

d As measured <strong>by</strong> CE.<br />

K1.74!10 K3<br />

K1.94!10 K3<br />

tumor tissues to examine the biodistribution and the toxicity. Then, gold nanocomposites with<br />

different surface charges and in different sizes were prepared and studied.<br />

The combined analytical results and properties of the PAMAM dendrimers used as templates<br />

are presented in Table 28.1 and Table 28.2. Details of template synthesis and characterization are<br />

described in the literature.<br />

Zeta potential measurements of {(Au 0 )9-PAMAM_E5.(NH2)110} in aqueous solutions<br />

confirmed that these nanoparticles are positively charged, indicating that the terminal amines<br />

can still be protonated after the formation of the nanocomposites. The net charge of amine-terminated<br />

dendrimers and their Au-nanocomposites was compared <strong>by</strong> PAGE and proved to be similar to their<br />

corresponding templates in the pHZ8.3 running buffer. 65<br />

TABLE 28.2<br />

Molecular Characteristics of Polycationic PAMAM Dendrimer Templates<br />

Mn a<br />

Mn b<br />

Dendrimer E3.NH2 E5.NH2<br />

6909 28,826<br />

6648 27250 c<br />

Diameter a (nm) 3.6 5.4<br />

Theoretical no. of NH2 groups a<br />

32 128<br />

Total no. of N ligands a<br />

62 254<br />

Charge/Mw ratio d (C/g) 8.973 !10 K3<br />

8.811 !10 K3<br />

Average no. of –NH2 groups b<br />

26.3 120 c<br />

Total average no. of –NZligands b<br />

54 238 c<br />

Charge/Mw ratio b (C/g) 8.12 !10 K3<br />

a<br />

Theoretical values.<br />

b<br />

Practical values determined from titration and GPC measurements.<br />

c<br />

Literature data.<br />

d<br />

Assuming theoretical values and full protonation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

8.73 !10 K3


564<br />

28.1.5.4 Synthesis of Tritium Labeled Dendrimers<br />

Earlier biodistribution studies of PAMAM dendrimers have been carried out with full generation<br />

PAMAM dendrimers that were radiolabeled <strong>by</strong> partially quaternizing the primary amine termini<br />

using 14 C, containing methyl iodide or 125 I labels. 17,78 Because these radiolabeled macromolecules<br />

carry permanent positive charges, this labeling procedure results in a mixture of products where the<br />

labeled molecules are chemically different from the unmodified population of dendrimers. The<br />

iodine-labeling procedure may lead to data where the distribution of radiolabeled materials is<br />

different from the overall distribution of the nanoparticles.<br />

Labeling of dendrimers <strong>by</strong> 3 H allows proper and accurate quantification of the PAMAM<br />

derivatives for the true in vivo biodistribution as a function of time because the labeled molecules<br />

are chemically identical with the unlabeled dendrimers. Regulation of surface charge with simultaneous<br />

labeling with tritium was achieved via acetylation <strong>by</strong> partially or fully reacting terminal<br />

primary amine groups of the PAMAM molecules. This synthetic step reduces the in vivo toxicity<br />

associated with the polyamine character (the terminal primary amine groups are positively charged<br />

at biologic pH values) and provides stable labels for the dendrimers. Increasing the degree of<br />

acetylation decreases the surface net positive charges; the fully acetylated generation five<br />

PAMAM displays about C5 mV zeta potential as opposed to C40 mV measured for the<br />

primary amine functional PAMAM_E5.NH2.<br />

The biodistribution of the fully acetylated dendrimer (NSD, Neutral Surface Dendrimer) and a<br />

partially acetylated one that has a partially positive surface charge have been compared. (The term<br />

of positively charged refers to the surface character in dilute aqueous solutions at biologic pH.)<br />

28.1.5.5 Folate-Targeted Devices<br />

Briefly, partially acetylated PAMAM was conjugated to folic acid as a targeting agent, and the<br />

remaining primary amines were capped using tritium labeled acetic anhydride. These conjugates<br />

were intravenously injected into immunodeficient mice bearing human KB tumors that overexpress<br />

the folic acid receptor. In contrast to non-targeted polymer, folate-conjugated nanoparticles<br />

concentrated in the tumor and liver tissue over fours days after administration. The tumor tissue<br />

localization of the folate-targeted polymer could be attenuated <strong>by</strong> prior i.v. injection of free folic<br />

acid. Using fluorescein labels instead of tritium labels, confocal microscopy was also used to<br />

confirm the internalization of the conjugates into the tumor cells. PAMAM dendrimers substituted<br />

with folic acid and methotrexate increased its anti-tumor activity and markedly decreased its<br />

toxicity, allowing therapeutic responses not possible with a free drug. 79<br />

28.1.5.6 Gold Composite Nanodevice Synthesis<br />

Nanotechnology for Cancer Therapy<br />

The {Au} gold/dendrimer composite nanodevices described below have been fabricated using four<br />

methods:<br />

1. By reactive encapsulation of gold in commercial (primary amine terminated and<br />

carboxylate terminated) PAMAM templates;<br />

2. By performing post-synthetic modificationonthesurfacegroupsofthe{Au}gold<br />

nanocomposites;<br />

3. By reactive encapsulation of gold in carboxylate functional PAMAM tecto-dendrimer<br />

templates; and<br />

4. By radiation polymerization of {Au} nanocomposites (Table 28.3).<br />

Method 28A1 fðAu 0 Þ 9:08KPAMAM_E5:ðNH 2Þ 120g C dZ5 nm, for short: fAuð0Þg C dZ5 nm, i.e., dZ<br />

5 nm amine surface gold/dendrimer nanocomposite (positively charged at pHZ7.4).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 565<br />

TABLE 28.3<br />

Summary of Nanodevices and Notations<br />

Description Full Annotation Acronym<br />

PSD PAMAM_E5:ðNH2Þ44ðNHCOCH3 Þ66 E5.(NH2)(NHOAc a )<br />

NSD PAMAM_E5:ðNHCOCH3 Þ110 E5.OAc a<br />

Targeted NSD PAMAM_E5:ðNHCOCH3 Þ104ðFAÞ6 E5.OAc a Neutral surface 5 nm gold CND fðAu<br />

-FA<br />

0Þ9:08KPAMAM_E5:ðNHCOCH3Þ120g C dZ5 nm fAuð0Þg 0 dZ5 nm<br />

Positive surface 5 nm gold CND fðAu0Þ9:08KPAMAM_E5:ðNH2Þ120g C dZ5 nm fAuð0Þg C dZ5 nm<br />

Negative surface 5 nm gold CND fðAu0Þ6:45KPAMAM_E4:5ðCOOHÞ60g K dZ5 nm fAuð0Þg K dZ5 nm<br />

Negative surface 11 nm gold CND fðAu 0 Þ60:09KðPAMAM_ððE5ÞðE3:ðCOOHÞ296ÞÞ Polymerized PCND b<br />

Polymerized PCND b<br />

The dZ5 nm positive nanodevices (Figure 28.7) were fabricated from PAMAM dendrimers<br />

and HAuCl4 as previously described. 5,36 Briefly, solution of HAuCl4 in methanol was added dropwise<br />

to a methanol solution of PAMAM dendrimer (MnZ26,000, functionalityZ120), resulting<br />

in a yellow solution of the tetrachloroaurate salt of the PAMAM dendrimer with a gold/dendrimer<br />

ratio of Au/DZ9.08), i.e., [PAMAM_E5.(NH2)120)(AuCl4)9.08]. The resulting yellowish solid was<br />

redissolved in water that then resulted in hydrolysis of the complex and led to nanocomposite<br />

formation (composite gold contentZ 6.44%). (Positive composite nanodevice,<br />

PCND)(Figure 28.9).<br />

Method 28A2 fðAu 0 Þ 6:45KPAMAM_E4:5ðCOOHÞ 60g K dZ5 nm, for short: fAuð0Þg K dZ5 nm, dZ5 nm<br />

carboxylate surface gold/dendrimer nanocomposite (negatively charged at pHZ7.4)<br />

Briefly, methanol solution of HAuCl4 was added drop-wise to an aqueous solution of generation<br />

4.5 PAMAM dendrimer (MnZ26,258, functionalityZ60), resulting in a red solution of the<br />

correspondent dendrimer nanocomposite (Au/DZ 6.45). (Composite gold contentZ4.41%)<br />

(Negative composite nanodevice, NgCND).<br />

Method 28B. Acetylation of the fAuð0Þg C dZ5 nm amine surface material, resulted in<br />

fðAu 0 Þ 9:08KPAMAM_E5:ðNHOCCH 3Þ 110g 0 dZ5 nm, for short: fAuð0Þg 0 dZ5 nm, an acetamide surface<br />

dZ5 nm gold/dendrimer nanocomposite that is nearly neutral at pHZ7.4 (neutral composite<br />

nanodevice, NCND).<br />

g K dZ11 nm<br />

fAuð0Þg K dZ11 nm<br />

fðAu0Þ5:7KE5:ðNH2Þg C 87; dZ22:2 nmand radioactive fðAu0Þ5:8g C dZ22 nm<br />

f 198 ðAu 0 ÞKE5:ðNH2Þg C 87; dZ22:2 nm<br />

f 198 ðAu 0 Þg C dZ22 nm<br />

fðAu0Þ5:7KE5:ðNH2ÞC 209; dZ29:7 nmgand fðAu0Þ5:7KE5:ðNH2Þg C f<br />

dZ29:7 nm<br />

198ðAu0ÞKE5:ðNH2Þg C 209; dZ29:7 nm f198ðAu0ÞKE5:ðNH2Þg C dZ29:7 nm<br />

PSD: Positive Surface Dendrimer; NSD: Neutral Surface Dendrimer; Ng: Negative; CND: Composite Nanodevice<br />

a Denotes radioactivity.<br />

b Numbers of primary particles have been calculated from volume, assuming full space-filling.<br />

Method 28C. The ðAu 0 Þ 60:09KðPAMAM_ððE5ÞðE3:ðCOOHÞ 296ÞÞ K<br />

dZ11 nm<br />

, for short:<br />

fAuð0Þg K dZ11 nm, i.e., the dZ11 nm carboxylate surface (negatively charged at pHZ7.4) gold/dendrimer<br />

nanocomposite was synthesized from PAMAM_E5(E3.COOH) 9, a succinamic acidterminated<br />

tecto-dendrimer (see Table 28.1). Briefly, methanol solution of HAuCl 4 was added<br />

drop-wise to an aqueous solution of PAMAM_E5(E3.COOH) 9 tecto-dendrimer (MnZ120,000,<br />

functionalityZ296), resulting in a red solution of the corresponding dendrimer nanocomposite<br />

(Au/DZ60.09; composite gold contentZ8.94%).<br />

Method 28D. Positively charged dO10 nm dendrimer nanocomposites were made from generation<br />

three and generation five templated nanocomposites <strong>by</strong> simultaneously exposing them to<br />

gamma and neutron radiation in a nuclear reactor above Tg, the glass transition temperature of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


566<br />

10 nm<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 28.9 TEM of single {(Au(0)nKE5 NH2} particles (marked with arrows) and various nanocomposite<br />

aggregates formed during sample preparation. (Reproduced from Balogh, L., Bielinska, A., Eichman, J. D.,<br />

Valluzzi, R., Lee, I., Baker, J. R., Lawrence, T. S., and Khan, M. K., Chim. Oggi (Chem. Today), 20 (5), 35–40,<br />

2002. With permission.)<br />

the polymers, (for PAMAMs TgZ20–35 C). Under these conditions, the neutrons activate the<br />

trapped gold metal clusters in the {(Au 0 ) 5.76} primary composite nanoparticles, and the gammaradiation<br />

cross links the organic network. Irradiation times were varied from 30 min to 3 h. Level of<br />

activation of 197 Au within the allowable degree of polymerization of the dendrimer host was<br />

controlled <strong>by</strong> setting the neutron irradiation time and limiting the gamma dose <strong>by</strong> led filters.<br />

Gold ( 197 Au) captures neutrons very efficiently because of its large cross-section. 198 Au (t1/2Z<br />

2.69 days) decays predominantly <strong>by</strong> beta-radiation (99%, 0.96 MeV) with a small gamma component<br />

(0.98%, 1.1 MeV) into a stable 198 Hg isotope. The build-up of activity for any given isotope<br />

is a function of several intrinsic factors, including isotopic abundance and neutron cross-section as<br />

well as a series of irradiation parameters, including neutron flux and energy spectrum. Therefore,<br />

although the general principles are valid for all the reactors, detailed experimental parameters of the<br />

procedure have to be independently worked out. Under these experimental conditions, efficiency<br />

was 50% of the target (theoretical) value (Figure 28.10).<br />

There were no changes observed in nanocomposite particle size at low doses. At higher<br />

irradiation doses, the size of nanoparticles increased and the particle size distribution became<br />

somewhat wider.<br />

The exact mechanism of cross linking is yet unknown. It is speculated that radical mechanisms<br />

that are responsible for beta-alanine cross linking (as well as local heat effects that are known to<br />

result in retro-Michael synthesis) may result in the rearrangement of the molecular structure. 80<br />

Data indicate that the product (similarly to silver/dendrimer nanocomposites) 76 is composed of<br />

primary composite nanoparticles and their aggregates (Table 28.4). Volume weighted averages are<br />

directly proportional to %w/w concentrations. Nevertheless, interaction with biologic objects (e.g.,<br />

cells) is determined <strong>by</strong> the number of interacting objects, there<strong>by</strong> necessitating determination of<br />

number averages as well. By number averages, each sample is predominantly composed of single<br />

particles (Figure 28.11).<br />

The size of the cross linked primary particles depends on the irradiation conditions; brief<br />

exposure to neutrons and gamma rays does not change particle size (Figure 28.12, left side), but<br />

it results in low activities. Longer irradiation activates the gold and simultaneously melts the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 567<br />

Measured activity microCi<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

Measd activity microCi<br />

Targeted activity microCi<br />

0<br />

0 5 10 15 20 25 30 35<br />

Irradiation time, min<br />

FIGURE 28.10 Comparison of targeted and measured specific activities of 198 Au as a function of irradiation<br />

time. (Reproduced from Balogh, L. P., Nigavekar, S. S., Cook, A. C., Minc, L., and Khan, M. K., PharmaChem<br />

2(4), 94–99, 2003. With permission.)<br />

dendrimers and cross links the primary particles into spherical and still soluble radioactive nanocomposites.<br />

Figure 28.12 shows the empirical correlation between particle diameter and irradiation<br />

time when pre-made {(Au 0 )6.53-PAMAM_E5.NH2} nanocomposites were treated in a reactor.<br />

In TEM, large particle structures that are probably responsible for the observed intense light<br />

scattering have been identified (Figure 28.13b). Visualization of these loose aggregates offers a few<br />

mechanistic clues. First, size and size distribution are unaffected when samples are irradiated with<br />

low doses in solution (Figure 28.13a). At higher doses, larger {Au} spherical particles form with a<br />

few superclusters present that are composed of primary {Au}n clusters loosely attached to each<br />

other <strong>by</strong> their surface.<br />

For approximate calculations, complete space-filling (100% packing) values for the cross<br />

linked primary particles and partial (50% packing) values for their aggregates were assumed.<br />

Based on volume averaged size data <strong>by</strong> DLS, a d Z 22.2G1.9 nm polymerized particle<br />

TABLE 28.4<br />

Particle Size Distributions of Soluble { 198 Au} C Nanocomposites After Irradiation of<br />

{(Au 0 )6.53-E5.NH2} (NICOMP Data)<br />

Sample {(Au 0 )6.53<br />

-E5.NH2}. A, B, C<br />

Mean<br />

Diam.<br />

(Nm)<br />

Volume Weighted Number Weighted<br />

Std. Dev.<br />

(Nm)<br />

Present<br />

(%)<br />

Mean<br />

Diam.<br />

(Nm)<br />

Std. Dev.<br />

(Nm)<br />

Present<br />

(%)<br />

Irradiation<br />

Time<br />

(Hr.)<br />

Zeta<br />

Potential<br />

(mV)<br />

fðAu0 C23:3 mV<br />

Þ6:53gnZ; dZ14:0G1:9 nm 14.0 1.9 99.4 13.4 1.4 100 1.0 C23.3<br />

Aggregate 453 45.1 0.6 0 0 0<br />

fðAu0 C22:7 mV<br />

Þ6:53gnZ; dZ16:6G1:7 nm 16.6 1.7 94.6 15.8 1.9 99.9 1.5 C22.7<br />

Aggregate 73.6 10.3 5.4 70.4 6.2 0.1<br />

fðAu0 C22:7 mV<br />

Þ6:53gnZ; dZ21:1G2:9 nm 21.1 2.9 91.2 20.2 2.2 99.9 2.0 C25.2<br />

Aggregate 111.2 17.2 8.8 104.9 4.0 0.1<br />

Data were measured after loss of radioactivity.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


568<br />

Rel %<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

LBCB-6-2 Normalized<br />

0<br />

10 28 81<br />

Diameter (nm)<br />

Num wt %<br />

Vol wt %<br />

FIGURE 28.11 Comparison of number weighted and volume weighted distributions reveals that although this<br />

nanocomposite solution contains a few clusters, each sample is predominantly composed of particles of similar<br />

size.<br />

fðAu 0 Þ C 5:76 g 87; dZ22:2 nmcontains 87 primary (dZ5 nm) DNC units and 499 Au 0 atoms per particle.<br />

A dZ29.7G2.9 nm polymerized gold/dendrimer nanocomposite particle ðfðAu 0 Þ C 5:69 g 209; dZ29:7 nmÞ<br />

contains 209 primary (dZ5 nm) DNC units and 1123 Au atoms. The aggregate of this secondary<br />

nanoparticles (with a packing fraction of 50%) would be composed of dZ22 nm particles (each<br />

containing 489 Au atoms in this example) that would suggest the presence of 51.4 {( 198 Au)n}<br />

Diameter. nm<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0 0.5 1 1.5 2 2.5<br />

Time of irradiation, h<br />

0<br />

FIGURE 28.12 Average diameter of Peak 1 in aqueous solutions of gold nanocomposites as a function of<br />

irradiation time (in equivalent position). Extensively irradiated nanocomposite samples became insoluble in<br />

water.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

0<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Num wt%


Dendrimer Nanocomposites for Cancer Therapy 569<br />

TABLE 28.5<br />

Particle Diameters of Radiation Polymerized Gold/Dendrimer Nanocomposites<br />

(Au/DZ5.69) According to Dynamic Light Scattering Data with NICOMP Data Analysis<br />

Sample<br />

Volume Weighted Number Weighted No. of Au atoms<br />

Mean<br />

Diameter<br />

(nm) Percent (%)<br />

Mean<br />

Diameter<br />

(nm) Percent (%) Per Host<br />

composite nanoparticles per aggregate (on average), containing 31,002 gold atoms in total. Data are<br />

shown in Table 28.5 and Table 28.6. The resulting nanocomposites preserved a net positive charge<br />

although measured zeta potential values (Table 28.6) were approximately half of what was<br />

measured for {(Au)56.61-PAMAM_E5.NH2} nanocomposites. 65 (The surface zeta potential for<br />

the latter nanocomposites was 41 mV as calibrated <strong>by</strong> dZ50 nm poly(styrene)sulfonate standards<br />

from Duke Scientific). Decrease of the surface positive charge indicates involvement of primary<br />

amine terminal groups in the polymerization reactions followed <strong>by</strong> their subsequent partial elimination<br />

from the host during radiation. The net positive zeta potential also supports the existence of<br />

the primary amine surface functions, explaining the formation of superclusters through<br />

surface attachment.<br />

28.1.5.7 Gold in the Nanocomposite Particles After Radiation Polymerization<br />

It is well known that the maximum of the surface plasmon resonances of noble metal nanocrystals<br />

in the UV–vis spectrum shifts toward longer wavelengths (from red to blue, i.e., toward lower<br />

energies) when particles aggregate. 81,82 However, the maxima in the UV–visible spectra of the<br />

{Au} nanocomposites are practically the same before and after radiation polymerization. The lack<br />

of shift indicates that the structural change is restricted to the host and the Au nanodomains remain<br />

isolated (Figure 28.14).<br />

28.1.5.8 Visualization of Gold/PAMAM Nanocomposite Uptake in Tumor Cells<br />

Gold mass<br />

Fraction (%)<br />

Primary particle: {(Au 0 )5.76}87 22.2 97.9 21.7 99.9 504 99.3<br />

Aggregate ({(Au 0 ) 5.76} 87) 12.5 51.5 2.1 51.4 0.1 3113 0.7<br />

In this demonstrative experiment, amine surface gold PAMAM nanocomposites were used to<br />

visualize {Au(0)}C in tumor tissue <strong>by</strong> TEM. One million B16F10 melanoma cells were<br />

TABLE 28.6<br />

NICOMP Size Distribution Data After Irradiation and Radioactive Decay for the Nanocomposite<br />

(Au/DZ5.76) Used in the Preliminary Therapeutic Experiments<br />

Sample<br />

Volume Weighted Number Weighted No. of Au atoms<br />

Mean<br />

Diameter<br />

(nm)<br />

Percent<br />

(%)<br />

Mean<br />

Diameter<br />

(nm)<br />

Percent<br />

(%) Per particle<br />

Mass<br />

Fraction (%)<br />

Primary particle: {(Au 0 )5.69}209 29.7 94.1 27.7 99.9 1123 97.5<br />

Aggregate ({(Au 0 )5.69}209)51.4 110.4 5.9 106.4 0.1 31,002 2.5<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


570<br />

25 nm<br />

(a) (b)<br />

20.51 nm<br />

27.82 nm<br />

20.16 nm<br />

205.47 nm<br />

FIGURE 28.13 (a) The TEM image illustrates the unperturbed size distribution of the {Au5.7-<br />

PAMAM_E5.NH2} gold nanocomposite when irradiated with low doses. (b) Almost identical dZ22 polymerized<br />

{Au}n particles forming a loose ({Au}n)m aggregate. (Reproduced from Balogh, L. P., Nigavekar, S.<br />

S., Cook, A. C., Minc, L., and Khan, M. K., PharmaChem 2(4), 94–99, 2003. With permission.)<br />

subcutaneously injected on the dorsal surface of a C57Bl6/J mouse and tumor allowed to develop<br />

for two weeks. The mice were then injected with the nanocomposite solution. Tumor and other<br />

organs (liver and lung tissue) were isolated and ultra-thin sections examined on a Philips CM100<br />

electron microscope. Figure 28.15 shows a TEM image of one of these sections with three tumor<br />

cells. The image indicates that 1.25 h after the direct injection only a few larger gold nanocomposite<br />

clusters are present in the interstitium. Most of the smaller clusters were found in the nucleus and<br />

cytoplasm. During the internalization of gold nanocomposites, diffusion through the cell wall<br />

Extinction<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

200<br />

300 400<br />

500 600 700<br />

Wavelength, nm<br />

Nanotechnology for Cancer Therapy<br />

26. nm<br />

{Au(0)} d=22 nm<br />

{Au(0)} d=11 nm<br />

{Au(0)} d=5 nm<br />

800 900 1000<br />

FIGURE 28.14 Comparison of UV–visible spectra of dZ5 nm (dashed line), dZ11 nm (dotted line), and<br />

dZ22 (continuous line) nm {Au(0)} gold nanocomposites.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 571<br />

FIGURE 28.15 Electron microscope image of mouse tumor tissue 1.25 h after {Au} injection. Unstained<br />

specimen, black dots represent gold nanocomposite clusters, dark areas within the nucleus are preferred for<br />

{Au}. (Reproduced from Balogh, L., Bielinska, A., Eichman, J. D., Valluzzi, R., Lee, I., Baker, J. R.,<br />

Lawrence, T. S., and Khan, M. K., Chim. Oggi (Chem. Today) 20(5), 35–40, 2002. With permission.)<br />

appears to be the dominant mechanism although transfer through a nuclear pore has also<br />

been observed.<br />

28.2 BIOLOGY<br />

As previously shown, nanocomposite dimensions are in the size range of small cellular machinery.<br />

The interactions that devices this small will have with biologic systems is truly unknown, and<br />

mathematical or other guiding principles to determine how exactly these small devices will behave<br />

in complex biologic systems that have potential interactions at the intracellular, cellular, multicellular,<br />

organ, and multi-organ system level are yet unknown. Research has demonstrated that until<br />

understanding of this process is had, it will be critical to empirically test these interactions as simple<br />

assumptions about nanodevice behavior in biologic systems are currently often incorrect.<br />

Interactions of nanoparticles with biologic objects should typically be surface interactions. In<br />

other words, they depend on the surface characteristics of both the biologic entity and the nanoparticle.<br />

This has been borne out in several of this team’s experiments and will be detailed below.<br />

The nanoparticle’s size is important, and this may be due to cellular or macromolecular interactions<br />

that are affected <strong>by</strong> size such as the ability to move through membranes, pores, membrane junctions,<br />

or recognition <strong>by</strong> certain cells (such as macrophages) for uptake. Charge of the nanoparticle is<br />

important because the Coulomb interactions are the strongest physical forces between any two<br />

objects. Other three-dimensional changes within the nanodevices that can affect their surface<br />

size/charge would also then be expected to affect biologic interactions.<br />

Biodistribution, or how these nanodevices move throughout complex biologic multi-organ<br />

systems, is a dynamic property and must be looked at as a function of time. Biodistribution can<br />

be affected <strong>by</strong> many factors, including site of injection or uptake, blood flow/rate, uptake resident<br />

time in the blood and various organs and organ systems, and clearance from the body (urine, feces,<br />

and sweat, for example) to name only a few of these factors. Nanodevices intravascularly injected<br />

will travel first to the heart then lungs, then back to the heart, then throughout the rest of the body.<br />

The heart and lung have first exposure to the nanodevices. Certain organ systems may have<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


572<br />

architecture (fenestrated structures for example) that may hasten or slow the time blood is exposed<br />

to the organ, the ability to enter organs, and the ability to stay in certain organs. Size/charge<br />

characteristics (particularly size) could have large influences on the route of clearance of these<br />

nanodevices. Detailed empiric biodistribution studies may eventually permit more detailed mathematical<br />

understanding of how different size/charge nanodevices will distribute over time in<br />

complex biologic systems.<br />

The vascular system is a critical organ system that needs detailed examination as it is the first<br />

organ system seen <strong>by</strong> intravascularly injected nanodevices and the main route of distribution to the<br />

rest of the body. Of particular importance to the study of cancer (and potentially other diseases such<br />

as macular degeneration, rheumatoid arthritis, etc.) is the angiogenic microvasculature. When these<br />

pathologic conditions occur or when a wound is healing, the normal existing microvasculature is<br />

stimulated to build a new microvasculature via budding off of the existing microvasculature. This<br />

process of neovascularization is called angiogenesis. It has potentially important implications for<br />

potential uses of nanodevices in pathologic and some non-pathologic biologic conditions as will be<br />

discussed below. A detailed understanding of the in vivo biodistribution of nanodevices will<br />

eventually permit the design of nanodevices that can specifically or selectively target specific<br />

organs or tissues in order to improve medical therapy.<br />

28.2.1 TARGETING AND NANODEVICE DELIVERY<br />

Nanotechnology for Cancer Therapy<br />

Targeting of tumor cells may be performed <strong>by</strong> either selective or specific ways as described in<br />

related literature. Direct receptor specific targeting is the one most often thought about, and<br />

nanocomposites permit the optimization of surface properties (via modulation of charge and attachment<br />

of targeting moieties) to maximally permit this direct targeting. Targeting groups can be<br />

conjugated to the host dendrimer’s surface 27 to allow the imaging agent to selectively bind to<br />

specific sites such as receptors on tumor cells to improve detection. For example, there are a<br />

number of cell surface receptors that are known to be over expressed on cancer cells, including<br />

epidermal growth factor receptor (EGFR) and the laminin receptor VLA-683. Nanocomposites<br />

could be constructed to specifically target these receptors.<br />

Nanodevices can also be constructed to target exposed receptors on the angiogenic microvasculature.<br />

Over the last few years, data have emerged supporting the concept that the<br />

microvasculatures of tissues and/or organs also have antigenic surface differences. Using a<br />

phage display technique, investigators have isolated peptide fragments that specifically bind to<br />

different tissue or organ microvasculatures in mice. 84–86 As part of this work, they also isolated<br />

peptides that specifically bind to angiogenic tumor microvasculature. 87–89 One of these angiogenic<br />

microvascular targeting peptides isolated is a three amino acid peptide, arginine–glycine–aspartate<br />

(RGD). RGD has been used to target chemotherapy and other agents to the angiogenic tumor<br />

microvasculature. 87,90,91 Radio-labeled linear or cyclic forms of the RGD peptide itself have<br />

been used to target the tumor microvasculature, permitting imaging with PET or SPECT. 92–97<br />

One could envision the use of RGD or other small molecules binding angiogenic microvasculature<br />

and present on the surface of the composite nanodevices to directly target the microvasculature.<br />

These studies are currently underway in this team’s laboratories.<br />

Non-target specific or non-selective targeting is another way to affect biodistribution of the<br />

nanodevices <strong>by</strong> not relying on the binding to a specific receptor but <strong>by</strong> taking advantage of the<br />

existing structural properties of complex biologic systems. It is well known that the angiogenic<br />

microvasculature is very different from the fully formed microvasculature. It is leakier than normal<br />

microvasculature, and there is also evidence that the tumor angiogenic microvasculature often lacks<br />

pericytes and has aberrant morphologies. 98–101 These differences between the tumor microvasculature<br />

and the normal microvasculature may result in enhanced permeability and retention effect<br />

(EPR), suggested first <strong>by</strong> Maeda. 102 The accumulation of macromolecules in the tumor was also<br />

found after i.v. injection of an albumin-dye complex (MwZ69,000) as well as after injection into<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 573<br />

normal and tumor tissues. The complex was retained only <strong>by</strong> tumor tissue for prolonged periods.<br />

With liposomes, the most effective size preferably retained in the tumor was suggested to be around<br />

50 nm. The Duncan group has also shown 103 that appropriately sized dendrimers and dendrimercis-platin<br />

complexes may become trapped in tumor vasculature as a result of the EPR effect. 104<br />

Using PAMAM dendrimers, it has also been demonstrated that polymers of different sizes have<br />

different permeability or cellular uptake, depending on the size (generation) and surface of the<br />

dendrimer (discussed in more detail below). For future therapy, these size and charge characteristics<br />

are exploited to specifically deliver nanoparticles through the tumor microvasculature<br />

and to the tumor or directly to the microvasculature itself using additional surface recognition to<br />

further improve avidity.<br />

28.2.1.1 Quantitative In Vivo Biodistribution of Dendrimer and Gold Composite<br />

Nanodevices (CND)<br />

As previously explained, composite nanodevices have several potential advantages over current<br />

technologies for the delivery or radiation or other therapies to tumors or the tumor microvasculature.<br />

In particular, even the smallest gold composite nanodevice can deliver greater than ten gold<br />

atoms per device to a target. This is a log fold more radiation than previously possible with the usual<br />

radioactive antibody techniques and offers a real potential to have greater than a log fold more<br />

radiation delivery per device than previously possible. The composite nanodevice architecture<br />

permits the simultaneous optimization of the carried inorganic component (gold or other inorganic<br />

material) and the dendrimer network (with or without receptor-specific targeting), providing the<br />

surface components of the fully formed nanodevice. To think of this clinical point of view, for<br />

example, targeting the nanodevice <strong>by</strong> the covalent linking of specific moieties to its surface and<br />

then carrying radioactive gold for imaging/therapy applications, the gold placement would not<br />

interfere with the targeter functioning on the surface. This optimization of the different compartments<br />

is a critical advantage of the composite nanodevices compared to other structures.<br />

28.2.1.2 Related Techniques in the Literature<br />

Illustrative of a few of these points is a study <strong>by</strong> Haifield et al. in the early 1990s. 55 This group<br />

carried out a partially successful attempt to increase the gold carrying content of antibodies <strong>by</strong><br />

covalently attaching gold clusters using sulfhydryl linkages to antibody fragments, and it then<br />

carried out biodistribution studies of these devices in mice. They were able to get some of the<br />

gold clusters averaging about 11 gold atoms to bind to various antibody fragments. The characterization<br />

of these clusters was done with TEM, and much more detailed characterization using several<br />

other techniques would be needed to be certain of the actual complexes formed. The size of the final<br />

assembly was not presented. They were able to make an antibody fragment-Au-cluster device that<br />

retained about 80% of its original binding activity (because of the interference of the gold clusters<br />

with the antibody fragment binding site). To stabilize the clusters, 21 phenyl groups were used, and<br />

this hydrophobic layer around the Au clusters caused increased uptake arguably in the blood cells.<br />

There was also very high uptake (trapping of clusters) in the kidney, probably as a result of the size<br />

of the antibody-Au cluster constructs. In both cases, the Au cluster chemistry/architecture interfered<br />

with the targeting component of the devices. More recent work on colloidal gold preparations <strong>by</strong><br />

Paciotti et al. 56 examined the in vivo biodistribution of a colloidal gold nanoparticle devices made<br />

of thiol–PEG (polyethylene glycol) and gold and recombinant human TNF (PT–cAu–TNF). The<br />

mechanism of the TNF linkage to the colloidal gold could not be delineated, but the preparations<br />

did appear to be 33 nm in size, and crude biodistribution measurements in a few organs showed<br />

relatively rapid clearance from the blood and deposition of the TNF in tumors. Much of the<br />

biodistribution of the nanoparticle was done <strong>by</strong> observing the color of organs as they turned<br />

more purple as they picked up the nanoparticle (possibly because of the formation of large gold<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


574<br />

nanoparticle aggregates that shifts the maximum of the plasmon peak toward blue). The data were<br />

not quantitative for whole particle determinations. The measurement of TNF in a few organs was<br />

quantitative. This nanoparticle appeared to have some interesting data supporting improved<br />

delivery to tumors compared to the non-PEGylated form of the nanodevice. It is unclear from<br />

the work, however, how this approach could be extended to permit other targeting with different<br />

agents or what improvements need to be made. There was no indication of how the Au–PEG–TNF<br />

components could be modified or optimized. Given the lack of knowledge of the actual chemistry<br />

holding the device together for the short few-hour time course of the experiment, it is unclear if this<br />

system can be modified or improved. These reports emphasize the advantages possible with the<br />

composite nanodevices, the need for much more detailed chemical characterization and biological<br />

understanding of how these devices distribute throughout the body, and how simple modifications<br />

may affect this biodistribution. 56,105<br />

Quantitative and detailed in vivo biodistribution data for dendrimer nanoparticles and, for the<br />

first time, composite nanodevices in tumor mouse model systems are presented below. Although<br />

there are many theories on how nanodevices might interact with complex animal systems, this team<br />

decided to empirically measure nanodevice biodistribution with several planned modifications of<br />

the nanodevices as a way to eventually begin to understand guiding principles of how nanodevices<br />

and, in particular, nanocomposites interact with organs/tissues. This data will eventually be used to<br />

develop a more detailed mathematical understanding of the process and to aid in the design of future<br />

nanodevices aimed at specific biologic understanding or therapy.<br />

As part of this approach, these groups carried out the first detailed and quantitative biodistribution<br />

studies on dendrimer nanodevices in tumor model systems. 77 They then expanded these studies<br />

to examine the first detailed and quantitative biodistribution studies of nanocomposites. 4 Important<br />

initial principles of the interactions of nanocomposites were determined, and predictions were made<br />

about how to examine nanodevices in general in complex biologic systems. 4,77,79<br />

28.2.1.3 The Instrumental Neutron Activation Analysis (INAA) Method<br />

Nanotechnology for Cancer Therapy<br />

Gold content of biologic samples <strong>by</strong> Au content of the nanodevices and tissue samples was<br />

determined <strong>by</strong> direct neutron irradiation. This approach was possible because (a) in the absence<br />

of oxygen, the gold/dendrimer nanocomposites are stable for a number of months (metal nanoclusters<br />

are effective oxygen-transfer catalysts), (b) PAMAMs do not break down in vivo because of<br />

enzymatic activity, and (c) the natural abundance of gold in mice and humans is close to zero.<br />

In the INAA technique, samples are activated <strong>by</strong> irradiation with neutrons, usually within the<br />

high flux fields of a nuclear reactor. Specific isotopes become radioactive <strong>by</strong> neutron capture-type<br />

nuclear reactions. After removal from the reactor, samples are allowed to decay (cool) to permit<br />

unwanted short-lived activity (for example, from Na) to diminish. Gamma-ray spectra from activated<br />

samples are then measured using solid-state detectors. 24 Specific isotopes may be identified<br />

<strong>by</strong> their gamma rays of characteristic energy, and quantitative determinations are made <strong>by</strong><br />

comparing peak areas in the spectrum to those of a standard reference material that has been<br />

irradiated and counted under identical conditions.<br />

Advantages of this analysis method are that (a) determination of CND biodistribution is very<br />

precise, (b) the biologic experiments can be performed on their own timescale and schedule,<br />

(samples can be prepared, accumulated, stored, and analyzed upon reactor availability), (c) high<br />

sensitivity may be achieved for gold, and (d) there is no biohazard material present after irradiation.<br />

However, it is a disadvantage that the dried and irradiated tissues are not suited for further study<br />

(TEM, pathology, immunology, etc.) The build-up of activity for a given isotope is a function of<br />

several intrinsic factors, including isotopic abundance and neutron cross-section as well as a series<br />

of irradiation parameters, including neutron flux and energy spectrum. Because irradiation parameters<br />

can vary slightly over time within a range of operating conditions, quantification of the<br />

activated isotope is rarely approached through direct calculation. Rather, elemental concentrations<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 575<br />

are determined through comparison on a weight-ratio basis from activities (measured as gamma<br />

counts-per-second) generated in a standard reference material of known composition, assuming all<br />

other parameters are held constant, including irradiation time, flux, decay time, counting time, and<br />

detector geometry.<br />

28.2.1.4 Determination of Gold Content Using INAA—Precision and Sensitivity<br />

All tissue samples were lyophilized and analyzed either at the Ford Nuclear Reactor of the Phoenix<br />

Memorial Laboratory (University of Michigan) or at the Oregon State University Radiation Center<br />

(OSU-RC, Corvallis, OR). A series of six gel check standards prepared with known gold concentrations<br />

(in 7% gelatin to mimic the organic tissue samples) were prepared to monitor measurement<br />

precision and sensitivity. Four replicates of a gold standard (NBS-SRM-2128-1, 1% Au <strong>by</strong> weight)<br />

prepared <strong>by</strong> pipetting approximately 100 mg of the liquid standard on to filter paper placed in<br />

sample vials were similarly weighed and desiccated. Samples and standards were irradiated for 2 h<br />

in core-face location 112B with a nominal thermal neutron flux of 10EC11. Following an initial<br />

2-day decay, a 15,000 s (live-time) count of gamma activity for Au-198 was recorded for each<br />

sample using a HPGe detector (33% relative efficiency) with a 4 00 counting geometry. Gold<br />

concentrations were determined through direct comparison on a weight ratio basis with the mean<br />

activity generated in four replicates of the Au standard. Sensitivities of !75 ppb were obtained,<br />

corresponding to 10 ng of Au (Figure 28.16). This method permits detailed and quantitative biodistribution<br />

measurements in mouse tumor model systems.<br />

28.2.1.5 Biodistribution of Positive and Neutral Surface Generation Five<br />

Dendrimers—Non-Specific Uptake<br />

In the first study of PAMAM dendrimers, the surface of PAMAM dendrimers were labeled with<br />

tritium 77 and studied the biodistribution of PAMAM_E5:ðNH 2Þ 44ðNHCOCH 3 Þ 66 a tritiated generation<br />

five (5 nm) positive surface dendrimer and PAMAM_E5:ðNHCOCH 3 Þ 110 neutral surface<br />

PAMAM dendrimers (PSD and NSD, respectively) in a mouse melanoma tumor model and in a<br />

prostate cancer mouse model. The feasibility of dendrimer delivery and subsequent quantification<br />

was also tested. Systemic administration of NSDs and PSDs via tail vein injection were carried out<br />

in tumor bearing mice. At different time points, the mice were euthanized and their organs<br />

Calculated Au, (ng)<br />

10 6<br />

10 5<br />

10 4<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.1<br />

y = 0.26864 * x ∧ (1.1488) R= 0.99971<br />

1 10 100 1000 10 4<br />

Measured Au, (ng)<br />

FIGURE 28.16 Comparison of Au measured <strong>by</strong> INAA and the theoretical (known) Au content measured in<br />

gel sample (gold concentrations were determined through direct comparison on a weight ratio basis with the<br />

mean activity generated in four replicates of the Au standard).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

10 5<br />

10 6


576<br />

harvested, weighed, and processed. The in vivo distribution of dendrimers was then determined via<br />

liquid scintillation counting of recovered tritium from the organs.<br />

Although neither particles localized selectively to tumor tissue, they did distribute to all organs<br />

tested and were recoverable within one hour post-injection. The vascular delivery of dendrimers<br />

could be homogenous in the organ (as seen in multiple samples of the liver analyzed). Both dendrimers<br />

rapidly cleared from blood, the % ID/g organ levels for PSD and NSD being 20.6 and 21.6 at<br />

5 min, 14.3 and 9.9 at 1 h, 1.7 and 0.9 at 1 day, 0.3 and 0.1 at 4 days, and 0.2 and 0.1 at 7 days postinjection,<br />

respectively. These organic nanoparticles localized to all major organs and tumor tissue<br />

and were rapidly cleared from the blood. Following an initial rapid clearance during the first day postinjection,<br />

concentrations were maintained at a relatively stable level in every tissue with a very slow<br />

decline over time. Lungs, kidneys, liver, and heart exhibited the highest uptake within an hour. Levels<br />

in the spleen and tumor followed this, and the negligible amounts were found in the brain tissue. More<br />

detailed tables and graphs of both PSD and NSD biodistributions are available in this team’s previous<br />

publication. 77 The PSD dendrimer nanoparticle biodistribution is shown in Figure 28.17. NSD<br />

biodistribution is very similar, but with lesser amounts throughout. In all organs and tissues tested<br />

(including tumor) the ratio of PSD:NSD was higher than 2:1, most likely because of increased<br />

retention of the positively charged dendrimers relative to neutral derivatives. It shows, in a quantitative<br />

manner, a global and differential effect of charge on organ and tissue biodistribution when the<br />

dendrimers are almost the same size (dZ5 nm).<br />

In another experiment, the nanoparticle biodistribution was determined in another tumor model<br />

system (human DU145 prostate cancer xenograft), and comparison was made to the previously<br />

determined B16 melanoma tumor model system (see Figure 28.17). The results are very similar in<br />

both models, supporting the general validity of the findings to multiple tumor and mouse types.<br />

In detailed excretion studies, it was shown that the dendrimers were excreted via both urine and<br />

feces (Figure 28.18). A total of 48.3% of NSD and 29.6% of PSD were excreted via urine over<br />

seven days. Within the first two hours, urinary NSD excretion was more than three-fold that of PSD,<br />

but at later sampling points, the daily excretion of both dendrimers was equivalent. This correlated<br />

with the increased retention of PSD seen in organs and tissues. Dendrimers were excreted via feces<br />

to a much lesser extent with a total of 5.41% of NSD and 2.96% of PSD recovered over seven days<br />

50<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

(a) Biodistribution of 5 nm PSD in C57BL6/J mice<br />

Blood<br />

Lungs<br />

Heart<br />

Liver<br />

Pancreas<br />

Time (h)<br />

Spleen<br />

Kidney<br />

Tumor<br />

1<br />

24<br />

96<br />

Brain<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Blood<br />

(b) Biodistribution of 5 nm PSD in nude mice<br />

Lungs<br />

Heart<br />

Liver<br />

Pancreas<br />

Time (h)<br />

FIGURE 28.17 Organ biodistribution of positive surface 5 nm nanoparticle (PSD) depicted as percent<br />

injected dose recovered per gram of organ (% ID/g). Panel (a) shows biodistribution in C57BL6/J mice<br />

subcutaneously carrying B16 melanoma tumors on the dorsal surface. Panel (b) shows the same nanoparticle<br />

injected into nude (nu/nu) mice carrying the human prostate carcinoma line DU145. nZ6 for panel (a) and<br />

nZ3 for panel (b).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

Spleen<br />

Kidney<br />

Tumor<br />

1<br />

24<br />

96<br />

Brain


Dendrimer Nanocomposites for Cancer Therapy 577<br />

25<br />

3<br />

20<br />

15<br />

NSD<br />

2<br />

10<br />

5<br />

PSD<br />

1<br />

0<br />

Time (Hours) 2 8 24 48 72 96 120 144 168<br />

0<br />

Time (Hours)<br />

% Injected dose<br />

NSD 24.92 12.82 6.29 2.76 0.72 0.33 0.18 0.14 0.15<br />

PSD 7.84 10.84 6.24 2.81 1.00 0.37 0.22 0.15 0.15<br />

(a) (b)<br />

(Figure 28.18). The dendrimers were mainly excreted via urine with a significant amount of urinary<br />

excretion occurring within 24 h post-injection. This was the first quantitative demonstration of the<br />

manner in which dendrimers are excreted from the body (urine) and of how this occurs over time.<br />

28.2.1.6 Biodistribution of 5 nm Gold Nanocomposites Compared with Dendrimer<br />

Templates<br />

As the next step in this research, this team examined composite nanodevices of differing surface. 4 It<br />

compared the biodistribution of 5 nm composite nanodevices with those of their respective<br />

dendrimer templates of the same surface charges and size. It found that although charge affected<br />

biodistribution of the same size nanodevices, surprisingly, the gold/dendrimer composite nanodevices<br />

had a very different organ/tissue biodistribution pattern in mice compared to the dendrimer<br />

nanoparticles of similar size and surface charge.<br />

28.2.1.7 Biologic Differences Between Dendrimers and Gold Composite Nanodevices<br />

This comparison is most accurate and particularly striking in the case of the acetamide (neutral)<br />

surface nanodevices where the size and surface charge are the same. There is significant and<br />

consistent high-level accumulation over time (selectivity) in the liver and spleen of mice intravenously<br />

injected with the gold nanocomposites (Figure 28.19a). In contrast, no such organ specific<br />

accumulation (selectivity) was seen with the NSD (Figure 28.19b).<br />

The overall (global) effects of surface charge are also different when the dendrimer nanoparticles<br />

are compared with the CNDs. With the dendrimers (Figure 28.19a and Figure 28.20a), and as<br />

discussed above, the partially positive surface macromolecules distributed into all organs at higher<br />

levels (about 2-fold higher levels) than that seen with the fully acetylated (neutral surface)<br />

dendrimer nanoparticles. 77 There were no significant organ-to-organ differences (or selectivity)<br />

seen. With the gold composite nanodevices, however, there are large organ-to-organ differences<br />

(selectivity) seen (Figure 28.19a and Figure 28.20b). There is significant spleen and liver accumulation<br />

(selectivity) seen with the fðAu 0 Þ 9:08KPAMAM_E5:ðNHCOCH 3Þ 120g C dZ5 nmneutral surface<br />

nanodevices over a prolonged time period. With the positive surface fðAu 0 Þ 9:08KPAMAM_E5:<br />

ðNH 2Þ 120g C dZ5 nmnanocomposite, there was significant accumulation noted in the kidney.<br />

Finally, a further major difference noted was that the unmodified primary amine terminated<br />

PAMAM dendrimer was previously shown to be toxic in vivo (causing the death of mice) as<br />

discussed in the introduction, and this is exactly why partially acetylated PAMAMs were used<br />

for in vivo studies as PSD. With the gold/dendrimer composite nanodevices, however, unmodified<br />

hosts were used to safely prepare respective nanocomposites for these mouse tumor model studies.<br />

Although the exact reasons are yet unknown, dendrimers and dendrimer nanocomposites are<br />

% Injected dose<br />

NSD<br />

PSD<br />

2 8 24 48 72 96 120 144 168<br />

NSD 0.40 2.63 1.43 0.74 0.03 0.03 0.05 0.05 0.05<br />

PSD 0.76 0.86 0.41 0.42 0.1 0.04 0.05 0.06 0.26<br />

FIGURE 28.18 Excretion of positive surface dendrimers (PSD) and neutral surface dendrimers (NSD) via<br />

urine and feces (a and b, respectively) at time points indicated. The data obtained from B16F10 melanoma<br />

mouse model (nZ5) are presented as a percentage of injected dose excreted per mouse.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


578<br />

Biodistribution of 5 nm Positive Au-Composite<br />

Biodistribution of 5 nm Partially Positive Dendrimer<br />

Nanodevice in the B16 mouse melanoma model<br />

Nanodevice in the B16 mouse melanoma model<br />

45<br />

45<br />

40 1 h<br />

40 1 h<br />

35 1 day<br />

35 1 day<br />

30 4 days<br />

30 4 days<br />

25<br />

25<br />

20<br />

20<br />

15<br />

15<br />

10<br />

10<br />

5<br />

5<br />

0<br />

BLD BRN PCS TMR HRT KDY LVR SPN LNG<br />

0<br />

BLD BRN PCS TMR HRT KDY LVR SPN LNG<br />

(a) (b)<br />

%ID/g organ<br />

FIGURE 28.19 Organ/tissue biodistribution of neutral surface Au-composite nanodevice or (Au0)9-PAMA-<br />

M_E5.(NHCOCH3)110} (a) and neutral surface dendrimer or PAMAM_E5.(NHCOCH3)110 (b) in tumor<br />

bearing mice. Abbreviations: BLD, blood; BRN, brain; PCS, pancreas, TMR, tumor; HRT, heart; KDY,<br />

kidney; LVR, liver; SPN, spleen; and LNG, lung. (The (b) has been reproduced from Nigavekar, S. S.,<br />

Sung, L. Y., Llanes, M., El-Jawahri, A., Lawrence, T. S., Becker, C. W., Balogh, L., and Khan, M. K.,<br />

Pharm. Res., 21(3), 476–483, 2004. With permission.)<br />

different enough to yield different biologic interactions (even though they are similar <strong>by</strong> charge and<br />

size in materials characterization), resulting in a different toxicity profile.<br />

The surface-to-volume ratio of nanodevices is much greater than for micron objects, enabling<br />

the surface to govern the interactions of nanoparticles with proteins, carbohydrates, or other macromolecules<br />

within the organs, tissues, or cells. It was previously believed that the surface of the 5 nm<br />

positive or NSDs would largely govern the biodistribution of the gold/dendrimer composite nanodevices.<br />

Even though gold nanocomposites of primary amine terminated PAMAMs typically form<br />

mixed structures, 41 the fact that the fðAu 0 Þ 9:08KPAMAM_E5:ðNHCOCH 3Þ 120g C dZ5 nm and<br />

fðAu 0 Þ 9:08KPAMAM_E5:ðNH 2Þ 120g C dZ5 nm nanodevices have very different biodistribution within<br />

the mouse tumor model system in comparison to the dendrimer templates indicates that, from a<br />

biologic point of view, these devices interact differently with macromolecules present on or in<br />

organs/tissues. A classic three-dimensional conformational change (as can be seen with proteins)<br />

and perhaps affecting these surface interactions cannot account for the difference in biodistribution<br />

%ID/g organ<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Biodistribution of 5 nm Neutral Au-Composite<br />

Nanodevice in the B16 mouse melanoma model<br />

1 h<br />

1 day<br />

4 days<br />

%ID/g organ<br />

%ID/g organ<br />

45<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Nanotechnology for Cancer Therapy<br />

Biodistribution of 5 nm Neutral Dendrimer<br />

Nanodevice in the B16 mouse melanoma model<br />

1 h<br />

1 day<br />

4 days<br />

BLD BRN PCS TMR HRT KDY LVR SPN LNG BLD BRN PCS TMR HRT KDY LVR SPN LNG<br />

(a) (b)<br />

FIGURE 28.20 Organ/tissue biodistribution of positive surface Au-composite nanodevice or {(Au0)9-<br />

PAMAM_E5.(NH 2) 110 (a) and partially positive surface dendrimer or PAMAM_E5.(NH 2)4 4 (NHCOCH 3) 66<br />

(b) in tumor bearing mice. Abbreviations: BLD, blood; BRN, brain; PCS, pancreas; TMR, tumor; HRT, heart;<br />

KDY, kidney; LVR, liver; SPN, spleen; and LNG, lung. (The (b) has been reproduced from Nigavekar, S. S.,<br />

Sung, L. Y., Llanes, M., El-Jawahri, A., Lawrence, T. S., Becker, C. W., Balogh, L., and Khan, M. K., Pharm.<br />

Res., 21(3), 476–483, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 579<br />

as all the studied particles are spherical. The increased rigidity and density of {Au(0)} composite<br />

nanodevices compared to the relative flexibility of the template dendrimers may explain some of<br />

this difference. Alternatively, perhaps another changed physiochemical property may partially be<br />

responsible for some differences in biodistribution. This aspect is presently under investigation.<br />

The results again indicate that those working on nanoparticles for biologic and medical uses<br />

need to understand in detail how nanodevices truly behave in vivo and that new nanodevices will<br />

have to be rigorously tested for biodistribution, toxicity, and other critical measures needed for the<br />

use of these devices in biologic systems. As future multifunctional nanodevices are made, it will be<br />

critical to carry out detailed biologic testing of these devices with respect to changes in charge, size,<br />

complexation, and surface substitution to gain an in depth understanding of how they will behave in<br />

these systems. Studies on dendrimers and composites have shown that this rigorous testing is<br />

critical and will be needed for other nanomaterials as well.<br />

28.2.1.8 The Importance of Size on Nanocomposite Biodistribution<br />

This team conducted a series of further studies that are currently in submission for publication<br />

where the effect of size and charge on biodistribution of nanocomposites is further examined. It<br />

examined 5, 11, and 22 nm nanocomposites with varying surface charge (partially positive,<br />

negative, and neutral). A key finding of these studies was that there are different organs that are<br />

selectively uptake charged nanoparticles, depending on the size and charge tested (see<br />

Figure 28.21). Importantly, when surface charge was kept constant and the size of the nanodevice<br />

was varied as with the 5 and 11 nm carboxylate (negative) surface nanocomposites, different<br />

nanodevice biodistribution was seen with differential organ selective uptake noted.<br />

Nanodevice levels in blood are higher for the fAuð0Þg K dZ5 nmgold nanocomposites than they are<br />

for the fAuð0Þg K dZ11 nm gold nanocomposites. Smaller size might preclude quick clearance from<br />

blood circulation, perhaps <strong>by</strong> uptake of the nanodevices into circulating blood cells. Nanodevice<br />

accumulation in the spleen appears to be determined <strong>by</strong> size as the fAuð0Þg K dZ11 nm negative surface<br />

gold nanocomposites show much higher levels of accumulation in the spleen than the 5 nm negative<br />

% ID/gorgan<br />

300<br />

250<br />

200<br />

150<br />

100<br />

50<br />

0<br />

5nm<br />

NgCND<br />

Biodistribution of 5 and 11 nm negative surface Au-composite<br />

nanodevices in the B16 mouse melanoma model<br />

11nm<br />

NgCND<br />

5nm<br />

NgCND<br />

11nm<br />

NgCND<br />

5nm<br />

NgCND<br />

11nm<br />

NgCND<br />

5nm<br />

NgCND<br />

11nm<br />

NgCND<br />

5nm<br />

NgCND<br />

Liver Spleen Lungs Kidneys Tumor<br />

Mouse organ<br />

FIGURE 28.21 Biodistribution of 5 and 11 nm negative surface Au-composite nanodevices in C57BL6/J<br />

mice bearing B16F10 melanoma. Shown here are nanodevice accumulation in tumor as well as in organs that<br />

show the highest accumulation of nanodevice that include the liver, spleen, lungs, and kidneys. Nanodevice<br />

levels were tested at times 5 min and hours 1, 24, and 96 post-injection.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

5 min<br />

1 h<br />

24 h<br />

96 h<br />

11nm<br />

NgCND


580<br />

surface Au-composite nanodevice. Liver accumulation was also consistently higher in the 11 nm<br />

nanocomposites. These studies indicate that size can affect distribution of the nanodevices with<br />

similar surface charges. This will have to be taken into consideration if this nanodevice were to be<br />

used to deliver radiation or toxic therapeutics, especially in the absence of a tumor targeting moiety<br />

on the surface of the nanodevice. This is the first set of data demonstrating the critical importance of<br />

the size of the nanodevices on their biodistribution in complex organisms.<br />

28.2.1.9 Biodistribution Summary<br />

Detailed quantitative biodistribution studies demonstrated the critical importance of surface size<br />

and charge on the interactions of nanocomposites with complex biologic systems. This team was<br />

able to empirically determine how differing size and charge affects this biodistribution, showing<br />

that selective organ uptake can be achieved with modulation of these parameters. It also showed<br />

that biodistribution profiles of Au-composite nanodevices were different from profiles of dendrimers<br />

of the same size and charge. These studies also revealed that an important principle of<br />

nanodevice design is to be fully aware of these parameters, including the importance of flexibility<br />

and density. Importantly, this team also demonstrated that it is difficult to predict what size/charge<br />

modification would produce a given biodistribution. The behaviors appear to be complex at this<br />

time. Completion of its ongoing research will provide a deeper insight into the biodistribution of<br />

nanoparticles of differing size and charge and create an important background to examine targeted<br />

devices. Without this detailed type of baseline study, there would be no way to ensure true targeting<br />

is occurring as envisioned targeting may be more a result of selective uptake because of size,<br />

charge, and other properties, and they are not due to the targeting moiety on the nanodevice.<br />

Understanding nanodevice behavior in complex animal systems will help to design the best nanodevices<br />

for targeting and suggest ways to protect organs from undesired effects. The detailed<br />

studies above will also allow researchers to begin a more detailed mathematical and pharmacokinetic<br />

analysis to explain nanodevice size/charge effects on biodistribution and permit them to make<br />

educated predictions as to their behavior in biologic systems in the future.<br />

28.2.2 TOXICITY STUDIES<br />

28.2.2.1 Long-Term Toxicity Studies of Tritiated Dendrimer Nanoparticles<br />

Nanotechnology for Cancer Therapy<br />

Initial long-term toxicity studies were carried out on both the dendrimer nanoparticles and on<br />

several of the composite nanodevices as indicated in Table 28.7. As in vitro toxicity data cannot<br />

yet be directly transposed to in vivo conditions, this team has carried out initial toxicity studies on<br />

the tritiated dendrimers. Because tumor-bearing mice would die as a result of their tumor-burden,<br />

long-term toxicity/biodistribution studies were carried out with normal C57BL/6J mice. PAMAM<br />

_E5:ðNH 2Þ 44ðNHCOCH 3 Þ 66 generation 5 PAMAM dendrimer derivative with a partially positive<br />

surface (PSD) and a neutral surface PAMAM_E5:ðNHCOCH 3 Þ 110 dendrimer (NSD) were evaluated<br />

for in vivo toxicity in 6–8 week old male C57BL6/J mice. The mice were observed for a period<br />

of 75 days. There were five mice in each group with a control group that was given PBS. Clinical<br />

toxicity tables were kept on each set of mice, and daily notations were made of any signs of<br />

weakness, lethargy, dehydration, change in mobility, or reflex. Weights of mice-given dendrimers<br />

did not differ from weights of mice in the control group (see Figure 28.22), and none of the mice<br />

died during observation. The labeled dendrimers were detectable in all organs tested (see Table II in<br />

Nigavekar et al. 2004). 77 This demonstrated that the nanoparticles could distribute to all tested<br />

organs and remain there at relatively stable levels for at least 12 weeks. The mice appeared healthy<br />

at that time as well, and no long-term clinical toxicity or weight loss was observed. The mice<br />

weights increased over the first 15 days and then were stably maintained with slight increment over<br />

the period observed. This team’s conclusion was that these dendrimers are non-toxic in this<br />

animal model.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 581<br />

TABLE 28.7<br />

Male C57BL6/J, 6–8-Week-Old Mice Were Injected with Tritiated Dendrimer Nanoparticles<br />

(5 nm PSD and NSD) and Au-Composite Nanodevices (5 nm PCND, 11 nm NgCND, and<br />

22 nm PCND) and The Were Monitored Over a Period of Time for Change in Weights<br />

Toxicity analysis<br />

Analysis of Mice Weights 5 (nm) PSD 5 nm NSD 5 nm PCND<br />

11 nm<br />

NgCND 22 nm PCND<br />

Time course of analysis<br />

(in days)<br />

Analysis of clinical<br />

75 75 75 49 49<br />

Toxicity O O O O O<br />

The mice were also monitored for signs of clinical toxicity such as weakness, lethargy, dehydration, change in mobility, or<br />

reflex. Death, as an end-point was noted as well.<br />

Weight (g)<br />

40<br />

30<br />

20<br />

10<br />

Weight (g)<br />

40<br />

30<br />

20<br />

10<br />

Weights of C57BL6/J<br />

mice given PSD<br />

0<br />

1 3 6 11 19 34 53 71<br />

Time (days)<br />

Weights of C57BL6/J<br />

mice given PBS<br />

0<br />

1 4 11 27 53 75<br />

Time (days)<br />

PSD 1<br />

PSD 2<br />

PSD 3<br />

PSD 4<br />

Weight (g)<br />

40<br />

30<br />

20<br />

10<br />

Control 1<br />

Control 2<br />

Control 3<br />

Control 4<br />

Control 5<br />

Weights of C57BL6/J<br />

mice given NSD<br />

0<br />

1 3 6 11 19 34 53 71<br />

Time (days)<br />

NSD 1<br />

NSD 2<br />

NSD 3<br />

NSD 4<br />

NSD 5<br />

FIGURE 28.22 Long-term monitoring of weights of non-tumor bearing C57BL/6J mice given (a) PBS as<br />

control (nZ5), (b) PSD (nZ4) and (c) NSD (nZ5). Individual weights of each mouse have been plotted as a<br />

function of time (days). PBS, phosphate buffered saline; PSD, positive surface dendrimer; NSD, neutral<br />

surface dendrimer. The notation PSD 1, for example, means the positive surface dendrimer injected into<br />

mouse 1.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


582<br />

28.2.2.2 Long-Term Toxicity of {Au(0)} Gold Composite Nanodevices<br />

Very limited data are available on the toxicity of noble metal dendrimer nanocomposites. Some<br />

preliminary data in cell culture systems provide some understanding of metal composite nanodevices<br />

in vitro. 5,76 It appears that toxicity is very similar to that of the dendrimer template, i.e., if the<br />

template is non-toxic, the noble metal composite nanoparticles are also non-toxic. 76<br />

To gain the insight into the in vivo toxicity of composite nanodevices, the team completed<br />

detailed in vivo toxicity analysis of fAuð0Þg C dZ5 nmthe 5 nm positively charged CND, fAuð0Þg K dZ11 nm<br />

the 11 nm negatively charged CND, and fðAu 0 Þ 5:8g C dZ22 nm the 22 nm positively charged CND in<br />

healthy mice. The mice were injected with the nanodevices as described above for the tritiated<br />

dendrimers. In brief, the 5 nm PCND, the 11 nm NgCND, and the 22 nm PCND were all evaluated<br />

for toxicity in 6–8 week old male C57BL6/J mice. Mice given the 5 nm nanodevice were observed<br />

for 75 days, whereas those given the 11 nm and 22 nm devices were observed for 49 days. There<br />

were five mice in each group with a control group that was given PBS. Daily toxicity tables were<br />

kept on each group of mice, and any signs of weakness, lethargy, dehydration, change in mobility,<br />

or reflexes noted. There was no clinical toxicity found. Weights of mice that were given dendrimers<br />

did not differ from weights of mice in the control group (see Figure 28.23), and none of the mice<br />

died during observation.<br />

28.2.2.3 Treatment of Tumors in Mouse Model Systems<br />

Reactive encapsulation permits the entrapment of different radionuclides within composite nanoparticles<br />

of defined size and targeting properties. An advantage of this approach would be the<br />

development of an image guided therapy where the radioactivity delivered to a tumor can be<br />

increased either <strong>by</strong> increasing the particle size or <strong>by</strong> increasing the number or specific activity of<br />

the guest atoms without destroying the targeting ability of the nanocomposite device. Additionally,<br />

Weight (g)<br />

Weight (g)<br />

Weights of C57BL6/J<br />

mice givenPBS<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

1 3 7 12 18 36 40 46<br />

Time (days)<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Weights of mice given<br />

5nm Au-PCND<br />

1 5 7 14 28 43 58 75<br />

Time (days)<br />

Control 1<br />

Control 2<br />

Control 3<br />

Control 4<br />

Control 5<br />

5nm PNCD 1<br />

5nm PNCD 2<br />

5nm PNCD 3<br />

5nm PNCD 4<br />

5nm PNCD 5<br />

Weight (g)<br />

Weight (g)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Nanotechnology for Cancer Therapy<br />

Weights of mice given<br />

11nm Au-NgCND<br />

1 3 7 12 18 36 40 46<br />

Time (days)<br />

Weights of mice given<br />

22nm Au-PCND<br />

1 3 7 12 18 36 40 46<br />

Time (days)<br />

NgNCD 1<br />

NgNCD 2<br />

NgNCD 3<br />

NgNCD 4<br />

NgNCD 5<br />

22nm PNCD 1<br />

22nm PNCD 2<br />

22nm PNCD 3<br />

22nm PNCD 4<br />

FIGURE 28.23 Long-term monitoring of weights of non-tumor bearing C57BL/6J mice given PBS (nZ5) as<br />

control (nZ5), 5 nm Au-PCND (nZ5), 22 nm Au-PCND (nZ4), and 11 nm Au-NgCND (nZ5). Individual<br />

weights of each mouse have been plotted as a function of time (days). Au-PCNDZgold positive surface<br />

composite nanodevice; Au-NgCNDZ gold negative surface composite nanodevice; PBSZ phosphate<br />

buffered saline. The notation PCND 1, for example, means that particular nanodevice injected into mouse 1.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 583<br />

varying the particular nature of the atoms (metals, isotopes) encapsulated in the CNDs will permit<br />

the application of different imaging techniques (e.g., film, SPECT, PET). Because these CNDs can<br />

encapsulate a great number (2–1024) 6,39 of radionuclides per macromolecule, they can deliver at<br />

least a log fold more radioactivity to tumors than is possible with current radioimmunotherapeutic<br />

technologies. This offers the added potential of imaging microscopic tumor burdens as seen in<br />

micrometastases or minimal residual disease states. In the future, this amplified signal and molecular<br />

targeting capability could greatly aid in advancing the field of molecular imaging. 106<br />

Significantly, these composite nanodevices could be used for a combined imaging and therapy<br />

of the tumors. 6,39,106<br />

Many attempts have been made to target radionuclides to tumors 107–114 with the best known<br />

approaches being the use of radiolabeled monoclonal antibody therapy or the use of surface<br />

chelators bound to dendrimers. A detailed review and critique of these approaches is beyond the<br />

scope of this chapter, and several useful reviews exist. 115–118 In the past, radiolabeled antibodies for<br />

direct tumor cell targeting have received wide interest for the purpose of receptor imaging. In<br />

general, however, the penetration of these macromolecules into tumor tissue has been problematic<br />

with typically only 0.001%–0.01% of the injected dose being localized to each gram of solid human<br />

tumor tissue in man. Several factors account for this poor penetration: antibodies in the circulation<br />

must travel across the endothelial cell layer and often through dense fibrous stroma before encountering<br />

tumor cells; the dense packing of tumor cells and tight junctions between epithelial tumor<br />

cells hinders transport of the antibody within the tumor mass; the absence of lymphatics within the<br />

tumor contributes to the buildup of a high interstitial pressure that opposes the influx of molecules<br />

into the tumor core; antibodies entering the tumor become absorbed to the perivascular regions <strong>by</strong><br />

the first tumor cells encountered, leaving none to reach tumor cells at sites farther from the blood<br />

vessels; and subpopulations of the rapidly mutating tumors can readily lose antigens targeted <strong>by</strong> the<br />

antibodies. Therefore, researchers have turned toward the development of small radiolabeled<br />

peptides that do not suffer from most of the aforementioned drawbacks. 107–118<br />

Tumor-directed antibody (or peptide) therapy has also not been able to deliver sufficient<br />

targeted dose because only one radioactive molecule can often be linked to a given antibody (or<br />

peptide), using a caged carrier to hold the radioactivity (e.g., DOTA). 119 The specific activity of the<br />

single radioactive moiety is also limited as it can damage the antibody if it is too high. This limits<br />

the detection of microscopic tumors, and higher doses leading to better detection have not been<br />

possible. This has also greatly limited any combination of imaging and therapy (where much higher<br />

dose delivery is needed). This derives at least partly from the fundamental problem that, to increase<br />

activity, one must increase substitution that tends to modify or obliterate the specificity of the<br />

antibody or targeting dendrimer. Finally, the ability to carry out different forms of imaging (PET or<br />

SPECT for example) is also limited <strong>by</strong> the chemistry necessary to place different isotopes onto the<br />

same targeting antibody. These fundamental problems can be overcome <strong>by</strong> the use of CNDs to<br />

deliver radioisotopes. 119<br />

Some of the fundamental problems that can be overcome <strong>by</strong> the use of targeted CNDs are those<br />

of solubility and dose delivery. First, many radioisotopes are often practically insoluble in water or<br />

body fluids, but they may become readily available in CNDs, i.e., in different sizes, contents, and<br />

surfaces. Second, use of CNDs permits at least a log-fold higher delivery of radioactivity than that<br />

available with current radioactive antibody therapies. As opposed to labeled antibodies that can<br />

carry only one radioisotope, 1–1000 nuclei can be encapsulated in one CND 6 , allowing the delivery<br />

of therapeutic dose amounts to tumor cells.<br />

To justify the targeted dose delivery <strong>by</strong> the any approach, it has to be first demonstrated that the<br />

treatment will be effective if the dose is delivered directly to the tumor (nanobrachytherapy). A<br />

proof-of-concept research of the therapeutic efficacy on { 198 Au} radioactive composite nanodevices<br />

in tumors (conceptually, other nuclides with the desired specific activity could also be used in<br />

the future, allowing the use of gamma, beta, alpha, or combinations of these radiations) has been<br />

conducted. To take advantage of the flexibility of composite nanodevice systems, Systemically<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


584<br />

Targeted RadioTherapy (StaRT) is being developed to target radioactive nanodevices either to<br />

tumors, to the tumor angiogenic microvasculature, or to both for the treatment of cancer. Before<br />

carrying out these targeted experiments, however, an idea of the level of doses that could be<br />

delivered <strong>by</strong> CNDs and if these doses were enough to slow or halt the growth of tumors was<br />

needed. In order to test this idea of nanobrachytherapy, specific doses of radioactive CNDs were<br />

directly injected into growing tumors, and the dose needed to significantly slow tumor growth was<br />

examined. Once this dose is known, calculating if targeted devices intravascularly injected could<br />

deliver these types of doses can be conducted (i.e., if it is feasible to proceed with targeted<br />

nanodevice delivery).<br />

B16 melanoma carrying mice were randomly sorted into three groups of seven mice with<br />

approximately similar tumor volumes (w500 mm 3 ). The control group was intra-tumorally<br />

injected with PBS buffer, whereas the remaining two groups were intra-tumorally injected with<br />

35 and 74 mCi of radioactive 198 Au carrying composite nanodevice, respectively. A second control<br />

experiment was carried out with the same nanodevice after the radioactivity of its guest component<br />

decayed over time. The now cold {Au(0)} nanocomposite (corresponding to the original 74 mCi<br />

sample) was injected intra-tumorally into one set of tumor bearing mice, and the second was again<br />

injected with PBS buffer as control. The tumor volumes of all mice were recorded every other day<br />

for eight days.<br />

When compared to tumors injected with either cold nanocomposite or PBS (untreated), there<br />

was a statistically significant slowing of tumor growth observed in tumor injected with 74 mCi of<br />

radioactive gold nanocomposite (pZ0.03 in both cases). Although injection with the 35 mCi CND<br />

did not significantly slow tumor growth, there was a noticeable trend toward slow tumor growth<br />

(see Figure 28.24). It should be possible with the use of fractionated experiments (multiple injections<br />

over time) to be able to safely deliver these types of doses to tumors using targeted devices.<br />

With radioactive antibodies, intraperitoneal/intravenous injections of 300–800 mCi of various<br />

isotope/antibody combinations have been successfully utilized to slow tumor growth in mouse<br />

xenograft models. 120–123 It was also assumed that about 5%–10% of the 800 mCi injected dose<br />

(approximately 40–80 mCi) to the tumor could be targeted as supported <strong>by</strong> data on folate-surfaced<br />

targeted dendrimers 79 and the literature above on other targeted devices. With these calculations,<br />

the 35–74 mCi needed for tumor delivery from intravascular injections are feasible, especially if<br />

fractionated treatment is contemplated.<br />

Tumor Volume (mm 3 )<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

(a) Radioactive Au Composite Nanodevice<br />

Treated Mice<br />

35 micro Ci<br />

74 micro Ci<br />

Untreated<br />

0 2 4 6 8<br />

Days<br />

Tumor Volume (mm 3 )<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

Nanotechnology for Cancer Therapy<br />

(b) Cold Au Composite Nanodevice<br />

Treated Mice<br />

COLD<br />

Untreated<br />

0 2 4 6 8<br />

Days<br />

FIGURE 28.24 Treatment of melanoma tumors using radioactive Au-composite nanodevice. B16F10 melanoma<br />

tumors were grown in C57BL6/J mice to sizes of w500 mm 3 . (a) The mice were intra-tumorally injected<br />

with PBS buffer (untreated control) and 35 and 74 mCi of radioactive 198 Au carrying composite nanodevice<br />

(CND), respectively (nZ7 for all three groups). (b) Mice were intra-tumorally injected with PBS buffer<br />

(untreated control) and cold nanodevice corresponding to the original 74 mCi radioactive Au-CND (nZ3<br />

for both groups). The tumor volumes of all mice were recorded every other day for eight days.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 585<br />

These experiments demonstrate for the first time that radiation therapy (nanobrachytherapy or<br />

systemic STaRT) with targeted composite nanodevices is feasible, and experiments are now being<br />

conducted to move this entire field forward with more detailed determination of the appropriate<br />

fractionation schemes for maximal tumor control using targeted composite nanodevices. Direct<br />

intratumoral injection gave a rough idea of the doses needed to do this, and they seemed reasonable<br />

given prior radioactive antibody data currently available in the literature. Importantly, for some<br />

tumors, direct injection of radioactive nanodevices may itself be a treatment.<br />

ACKNOWLEDGMENTS<br />

This research has been funded with federal funds in part from the U.S. Department of Energy under<br />

Award No. FG01-00NE22943 in part from the National Cancer Institute and National Institutes of<br />

Health, under Contract No. NO1-CO-97111 and RO1 CA104479 and in part <strong>by</strong> the Department of<br />

Defense Grant# DAMD17-03-1-0018. Silver nanocomposite research was sponsored <strong>by</strong> the Army<br />

Research Laboratory under the contract of DAAL-01-1996-02-0044. Thanks to D. Sorenson and C.<br />

Edwards of the Microscope and Imaging Laboratory at the University of Michigan for their<br />

technical support, to Leah Minc OSU-RC for INAA analysis, to Shraddha Nigavekar, Bindu<br />

Nair, Muhammed Taju Kariapper, Lok Yun Sung, Mikel Llanes, Kerstin May, Areej El-Jawari,<br />

Alla Kwitney, and Amber Warnat for their significant contributions on various components of the<br />

biologic components of these projects, and to Wojciech Lesniak, Kai Sun, X. Shi, and the<br />

University of Michigan Electron Microbeam Analysis Laboratory for their contributions to the<br />

synthesis and characterization components of these studies.<br />

DEFINITIONS<br />

Nanoscience The science of the phenomena related to nanoscale materials and/or devices. It refers<br />

to the multi-disciplinary study of materials, processes, and mechanisms that have distinctive<br />

properties (chemical, physical, and biological) from the molecular (i.e., !1 nm) to the<br />

sub-micrometer scales (i.e., O100 nm) because of quantum-size effects.<br />

Nanotechnology A wide range of possible technologies that encompass the incorporation of<br />

engineered or manufactured devices or materials with critical length scales between<br />

approximately 1 nm and 100 nm, and applications that exploit specific properties and<br />

functions of nanosized materials and/or devices.<br />

Particle Any object that has a persistent surface and an interior that may be different from<br />

the surface.<br />

Nanodevice A surface modified multifunctional nanoparticle that is able to perform<br />

different functions.<br />

Cluster Aggregations of atoms that are too large to be referred to as molecules and too small to<br />

resemble small pieces of crystals. Clusters generally do not have the same structure or<br />

atomic arrangement as a bulk solid and can change structure with the addition of just one<br />

or a few atoms. As the number of atoms (and their degree of freedom) increases, eventually<br />

a crystal-like structure may be established. Cluster science is devoted to understanding the<br />

changes in fundamental properties of materials, as a function of size, evolving from<br />

isolated atoms or small molecules to a bulk phase. 1,2<br />

Aggregates Dendrimers or nanoparticles held together <strong>by</strong> weak physical forces.<br />

Nomenclature To describe the complex structure of these nanoscopic complexes and nanocomposites,<br />

the following convention is used to describe the composition of the synthesized<br />

nanomaterials: brackets denote complexes, and braces represent nanocomposite<br />

structures. Within the brackets or braces, the complexed or encapsulated components<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


586<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

are listed (with an index of their average number per dendrimer molecule) followed <strong>by</strong> the<br />

family of dendrimers and the terms used for identification, i.e., core, generation, and<br />

surface. Naming of the materials will follow this pattern. For instance,<br />

PPI_DAB4.(NH 2) 16 describes a diaminobutane core poly(propyleneimine) dendrimer<br />

molecule with 16 primary amine terminal groups, and PAMAM_E5.(NH2)44(NHOAc)66<br />

denotes a generation five ethylenediamine core poly(amidoamine) dendrimer material<br />

with an average of 44 primary amine and 66 acetamide terminal groups (Please observe<br />

that the sum of these numbers are less than the theoretical 128, indicating that these are<br />

measured average numbers). All known properties can be listed <strong>by</strong> using this principle.<br />

Nanocomposites are also named according to the same pattern {(component#1)i(component#2)j.-FAMILY_Core.Generation.Terminal<br />

group}Therefore, the formula<br />

{(Au 0 ) 14.5-PAMAM_E4.NH 2} denotes a gold dendrimer nanocomposite where a generation<br />

4 amine terminated ethylenediamine (EDA or E for short) core poly(amidoamine)<br />

(PAMAM) dendrimer contains 14.5 zerovalent gold atoms per dendrimer molecule on<br />

average. The above scheme provides a relatively simple, but consistent, way to identify<br />

materials with a complicated structure. Additional descriptors may be added to the superscripts/superscripts<br />

and parentheses can be embedded into each other, e.g.,<br />

fðAu 0 Þ10-E5:NH2g250G10 The composite listed above denotes an aggregate particle composed of 250G10 primary<br />

composite nanoparticles. When it is trivial, that dendrimer family is used; one may follow<br />

the shorter pattern and omit naming the family. Therefore, {(Au 0 )10-PAMAM_E5.NH2}<br />

and {(Au 0 ) 10-E5.NH2} may be considered to be equivalent names if only PAMAMs<br />

are used.<br />

1. Alivisatos, A. P. and Schultz, P. G., Organization of ‘nanocrystal molecules’ using DNA, Nature,<br />

382 (6592), 609–611, 1996.<br />

2. Schmid, G., Large clusters and colloids. Metals in the embryonic state, Chem. Rev., 92 (8),<br />

1709–1727, 1992.<br />

3. Jemal, A., Tiwari, R. C., Murray, T., Ghafoor, A., Samuels, A., Ward, E., Feuer, E. J., and Thun,<br />

M. J., Cancer statistics, CA Cancer J. Clin., 54 (1), 8–29, 2004.<br />

4. Khan, M. K., Nigavekar, S. S., Minc, L. D., Kariapper, M. S., Nair, B. M., Lesniak, W. G., and<br />

Balogh, L. P., In vivo biodistribution of dendrimers and dendrimer nanocomposites—implications<br />

for cancer imaging and therapy, Technol. Cancer Res. Treat., 4 (6), 603–613, 2005.<br />

5. Bielinska, A., Eichman, J. D., Lee, I., Baker, J. R., and Balogh, L., Imaging {Au0-PAMAM} gold–<br />

dendrimer nanocomposites in cells, J. Nanopart. Res., 4, 395–403, 2002.<br />

6. Grohn, F., Bauer, B. J., Akpalu, Y. A., Jackson, C. L., and Amis, E. J., Dendrimer templates for the<br />

formation of gold nanoclusters, Macromolecules, 33 (16), 6042–6050, 2000.<br />

7. McDevitt, M. R., Ma, D., Lai, L. T., Simon, J., Borchardt, P., Frank, R. K., Wu, K., Pellegrini, V.,<br />

Curcio, M. J., and Miederer, M., et al., Tumor therapy with targeted atomic nanogenerators, Science,<br />

294 (5546), 1537–1540, 2001.<br />

8. Newkome, G. R., Moorefield, C. N., and Vögtle, F., Eds., Dendritic Macromolecules: Concepts,<br />

Syntheses, Perspectives, Wiley-VCH, New York, 1996.<br />

9. Frechet, J. M. J. and Tomalia, D. A., Eds., Dendrimers and Other Dendritic Polymers, Wiley, New<br />

York, p. 647, 2001.<br />

10. Mansfield, M. L. and Klushin, L. I., Monte Carlo studies of dendrimer macromolecules, Macromolecules,<br />

26 (16), 4262–4268, 1993.<br />

11. Karatasos, K., Adolf, D. B., and Davies, G. R., Statics and dynamics of model dendrimers as studied<br />

<strong>by</strong> molecular dynamics simulation, J. Chem. Phys., 115, 5310–5318, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 587<br />

12. Murat, M. and Grest, G. S., Molecular dynamics study of dendrimer molecules in solvents of varying<br />

quality, Macromolecules, 29 (4), 1278–1285, 1996.<br />

13. Welch, P. and Muthukumar, M., Tuning the density profile of dendritic polyelectrolytes, Macromolecules,<br />

31 (17), 5892–5897, 1998.<br />

14. Tomalia, D. A., Baker, H., Dewald, J., Hall, M., Kallos, G., Martin, S., Roeck, J., Ryder, J., and<br />

Smith, P., Polym. J. (Tokyo), 17 (1), 117–132, 1985.<br />

15. Grayson, S. M. and Fréchet, J. M. J., Convergent dendrons and dendrimers: From synthesis to<br />

applications, Chem. Rev., 101 (12), 3819–3868, 2001.<br />

16. Duncan, R. and Izzo, L., Dendrimer biocompatibility and toxicity, Adv. Drug Deliv. Rev., 57 (15),<br />

2215–2237, 2005.<br />

17. Roberts, J. C., Bhalgat, M. K., and Zera, R. T., Preliminary biological evaluation of polyamidoamine<br />

(PAMAM) starburst dendrimers, J. Biomed. Mater. Res., 30 (1), 53–65, 1996.<br />

18. Tomalia, D. A., Dewald, J. R., Hall, M. J., Martin, S.J., and Smith, P. B., First SPSJ International<br />

Polymer Conference, Kyoto, Japan, p. 65, 1984.<br />

19. Tomalia, D. A., Naylor, A. M., and Goddard, W. A., Starburst dendrimers: Molecular-level control<br />

of size, shape, surface chemistry, topology, and flexibility from atoms to macroscopic matter,<br />

Angew. Chem. Int. Ed. Engl., 29 (2), 138–175, 1990.<br />

20. D’Emanuele, A. and Attwood, D., Dendrimer–drug interactions, Adv. Drug Deliv. Rev., 57 (15),<br />

2147–2162, 2005.<br />

21. Svenson, S. and Tomalia, D. A., Dendrimers in biomedical applications—reflections on the field,<br />

Adv. Drug. Deliv. Rev., 57 (15), 2106–2129, 2005.<br />

22. Tomalia, D. A. and Dvornic, P. R., In Polymeric Materials Encyclopedia, Salamone, J. C., Ed., Vol.<br />

3, CRC Press, Boca Raton, FL, pp. 1814–1830, 1996.<br />

23. Bielinska, A., Kukowska-Latallo, J. F., Johnson, J., Tomalia, D. A., and Baker, J. R. Jr., Regulation<br />

of in vitro gene expression using antisense oligonucleotides or antisense expression plasmids transfected<br />

using starburst PAMAM dendrimers, Nucleic Acids Res., 24 (11), 2176–2182, 1996.<br />

24. El-Sayed, M., Kiani, M. F., Naimark, M. D., Hikal, A. H., and Ghandehari, H., Extravasation of<br />

poly(amidoamine) (PAMAM) dendrimers across microvascular network endothelium, Pharm. Res.,<br />

18 (1), 23–28, 2001.<br />

25. Malik, N., Evagorou, E. G., and Duncan, R., Dendrimer-platinate: A novel approach to cancer<br />

chemotherapy, Anticancer Drugs, 10 (8), 767–776, 1999.<br />

26. Raduchel, B., Schmitt-Willich, S., Ebert, J., Frezel, T., Misselwitz, B., and Weinmann, H. J.,<br />

Synthesis and Characterization of Novel Dendrimer-Based Gadolinium Complexes as MRI Contrast<br />

Agents for the Vascular System, American Chemical Society, Washington, DC, p. 278, 1998.<br />

27. Wilbur, D. S., Pathare, P. M., Hamlin, D. K., Buhler, K. R., and Vessella, R. L., Biotin reagents for<br />

antibody pretargeting. 3. Synthesis, radioiodination, and evaluation of biotinylated starburst dendrimers,<br />

Bioconjug. Chem., 9 (6), 813–825, 1998.<br />

28. Shi, X., Banyai, I., Islam, M. T., Lesniak, W., Davis, D. Z., Baker, J. R. J., and Balogh, L. P.,<br />

Generational, skeletal and substitutional diversities in generation one poly(amidoamine) dendrimers,<br />

Polymer, 46, 3022–3034, 2005.<br />

29. Uppuluri, S., Swanson, D. R., Brothers, H. M., Piehler, L. T., Li, J., Meier, D. J., Hagnauer, G. L., and<br />

Tomalia, D. A., Tecto(dendrimer) core-shell molecules: Macromolecular tectonics for the systemic<br />

synthesis of larger controlled structure molecules, Polym. Mater. Sci. Eng., 80, 55–56, 1999.<br />

30. Uppuluri, S., Swanson, D. R., Brothers, H. M. I., Piehler, L. T., Li, J., Meier, D. J., H., G. L., Tomalia,<br />

D.A, Core-Shell Tecto(Dendrimers): A New Class of Regio-Specifically Cross-Linked Polymers,<br />

Polymers for Advanced Technologies (PAT), Tokyo, Japan, August 31–September 5, pp. 94–95, 1999.<br />

31. Quintana, A., Raczka, E., Piehler, L., Lee, I., Myc, A., Majoros, I., Patri, A. K., Thomas, T., Mule, J.,<br />

and Baker, J. R. Jr., Design and function of a dendrimer-based therapeutic nanodevice targeted to<br />

tumor cells through the folate receptor, Pharm. Res., 19 (9), 1310–1316, 2002.<br />

32. Balogh, L., Swanson, D.R., Spindler, R., Tomalia., D.A., Formation and characterization of<br />

dendrimer-based water soluble inorganic nanocomposites, In Proceedings of the American Chemical<br />

Society, Division of Polymeric Materials Science and Engineering, Vol. 77, Los Vegas, VV, pp. 118–<br />

119, 1997.<br />

33. Tan, N. B., Balogh, L., and Trevino, S., Structure of metallo-organic nanocomposites produced from<br />

dendrimer complexes, Proc. ACS Polym. Mater.: Sci. Eng., 77, 120, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


588<br />

Nanotechnology for Cancer Therapy<br />

34. Balogh, L. and Tomalia, D. A., Dendrimer-templated nanocomposites. I. Synthesis of zero-valent<br />

copper nanoclusters, J. Am. Chem. Soc., 120, 7355–7356, 1998.<br />

35. Zhao, M., Sun, L., and Crooks, R. M., Preparation of Cu nanoclusters within dendrimer templates,<br />

J. Am. Chem. Soc., 120 (19), 4877–4878, 1998.<br />

36. Esumi, K., Suzuki, A., Aihara, N., Usui, K., and Torigoe, K., Preparation of gold colloids with UV<br />

irradiation using dendrimers as stabilizer, Langmuir, 14 (12), 3157–3159, 1998.<br />

37. Larré, C., Donnadieu, B., Caminade, A. M., and Majoral, J. P., Regioselective gold complexation<br />

within the cascade structure of phosphorus-containing dendrimers, Chem. Eur. J., 4 (10), 2031–2036,<br />

1998.<br />

38. Bosman, A. W., Janssen, H. M., and Meijer, E. W., About dendrimers: Structure, physical properties,<br />

and applications, Chem. Rev., 99, 1665–1688, 1999.<br />

39. Diallo, M., Balogh, L., Shafagati, A., Johnson, J. H. J., Goddard, W. A. I., and Tomalia, D. A., 2-<br />

Poly(amidoamine) dendrimers: A new class of high capacity chelating agents for Cu (II) ions,<br />

Environ. Sci. Technol., 33 (5), 820–824, 1999.<br />

40. Diallo, M. S., Christie, S., Swaminathan, P., Balogh, L., Shi, X., Um, W., Papelis, C., Goddard,<br />

W. A. III, and Johnson, J. H. Jr., Dendritic chelating agents. 1. Cu(II) binding to ethylene diamine<br />

core poly(amidoamine) dendrimers in aqueous solutions, Langmuir, 20 (7), 2640–2651, 2004.<br />

41. Balogh, L., Valuzzi, R., Laverdure, K. S., Gido, S. P., Hagnauer, G. L., and Tomalia, D. A.,<br />

Formation of silver and gold dendrimer nanocomposites, J. Nanopart. Res., 1 (3), 353–368, 1999.<br />

42. Balogh, L., Laverdure, K. S. Gido, S. P., Mott A. G., Miller, M. J., Ketchel, B. P., Tomalia, D. A.<br />

Dendrimer–metal nanocomposites, In Organic/Inorganic Hybrid Materials, pp. 69–75, 1999.<br />

43. Hong, S., Bielinska, A. U., Mecke, A., Keszler, B., Beals, J. L., Shi, X., Balogh, L., Orr, B. G., Baker,<br />

J. R. Jr., and Banaszak Holl, M. M., Interaction of poly(amidoamine) dendrimers with supported<br />

lipid bilayers and cells: Hole formation and the relation to transport, Bioconjug. Chem., 15 (4),<br />

774–782, 2004.<br />

44. Martin, C. R., Science, 266, 1961–1966, 1994.<br />

45. Shaw, C. F. III, Gold-based therapeutic agents, Chem. Rev., 99, 2589–2600, 1999.<br />

46. El-Sayed, I. H., Huang, X., and El-Sayed, M. A, Selective laser photo-thermal therapy of epithelial<br />

carcinoma using anti-EGFR antibody conjugated gold nanoparticles, Cancer Lett., 2005.<br />

47. Hirsch, L. R., Stafford, R. J., Bankson, J. A., Sershen, S. R., Rivera, B., Price, R. E., Hazle, N. J., and<br />

West, J. L., Nanoshell-mediated near-infrared thermal therapy of tumors under magnetic resonance<br />

guidance, Proc. Natl Acad. Sci. U.S.A., 100 (23), 13549–13554, 2003.<br />

48. Loo, C., Lowery, A., Halas, N., West, J., and Drezek, R., Immunotargeted nanoshells for integrated<br />

cancer imaging and therapy, Nano Lett., 5 (4), 709–711, 2005.<br />

49. Loo, C., Lin, A., Hirsch, L., Lee, M., Barton, J., Halas, N., West, J., and Drezek, R., Nanoshell-enabled<br />

photonics-based imaging and therapy of cancer, Technol. Cancer Res. Treat., 3, 33–40, 2004.<br />

50. Holm, H. H. III, The history of interstitial brachytherapy of prostatic cancer, Semin. Surg. Oncol., 13,<br />

431–437, 1997.<br />

51. Root, S. W., Andrews, G. A., Knieseley, R. M., and Tyor, M. P., The distribution and radiation<br />

effects of intravenously administered colloidal Au 198 in man, Cancer, 7, 856–866, 1954.<br />

52. Rubin, P. and Levitt, S. H., The response of disseminated reticulum cell sarcoma to the intravenous<br />

injection of colloidal radioactive gold, J. Nucl. Med., 5, 581–594, 1964.<br />

53. Flocks, R. H., Kerr, H. D., Elkins, H. B., and Culp, A., Treatment of carcinoma of the prostate <strong>by</strong><br />

interstitial radiation with radio-active gold: A preliminary report, J. Urol., 68, 510–522, 1952.<br />

54. Loening, S. A., Gold seed implantation in prostate brachytherapy, Semin. Surg. Oncol., 13 (6),<br />

419–424, 1997.<br />

55. Hainfeld, J. F., Foley, C. J., Srivastava, S. C., Mausner, L. F., Feng, N. I., Meinken, G. E., and<br />

Steplewski, Z., Radioactive gold cluster immunoconjugates: Potential agents for cancer therapy, Int.<br />

J. Radiat. Appl. Instrum. Part B, Nucl. Med. Biol. (Oxford), 17 (3), 287–294, 1990.<br />

56. Paciotti, G. F., Myer, L., Weinreich, D., Goia, D., Pavel, N., McLaughlin, R. E., and Tamarkin, L.,<br />

Colloidal gold: A novel nanoparticle vector for tumor directed drug delivery, Drug Deliv., 11 (3),<br />

169–183, 2004.<br />

57. Garcia, M. E., Baker, L. A., and Crooks, R. M., Preparation and characterization of dendrimer–gold<br />

colloid nanocomposites, Anal. Chem., 71 (1), 256–258, 1999.<br />

58. Tomalia, D. A., Balogh, L. P., US Patent No. 6,664,315B2, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 589<br />

59. Balogh, L., Tomalia, D. A., and Hagnauer, G. L., Revolution of nanoscale proportions, Chem.<br />

Innovation, 30 (3), 19–26, 2000.<br />

60. Ottaviani, M. F., Valluzzi, R., and Balogh, L., Internal structure of silver-poly(amidoamine)<br />

dendrimer complexes and nanocomposites, Macromolecules, 35 (13), 5105–5115, 2002.<br />

61. Balogh, L., Bielinska, A., Eichman, J. D., Valluzzi, R., Lee, I., Baker, J. R., Lawrence, T. S., and Khan,<br />

M. K., Dendrimer nanocomposites in medicine, Chim. Oggi (Chem. Today), 20 (5), 35–40, 2002.<br />

62. Cotton F. A., Willinson G., Advanced Inorganic Chemistry, 4 ed., Wiley, New York, pp. 1052–1055,<br />

1980.<br />

63. Lee, W. I., Bae, Y., and Bard, A. J., Strong blue photoluminescence and ECL from OH-terminated<br />

PAMAM dendrimers in the absence of gold nanoparticles, J. Am. Chem. Soc., 126, 8358–8359, 2004.<br />

64. Esumi, K., Suzuki, A., Yamahira, A., and Torigoe, K., Role of poly(amidoamine) dendrimers for<br />

preparing nanoparticles of gold, platinum, and silver, Langmuir, 16 (6), 2604–2608, 2000.<br />

65. Balogh, L., Ganser, T. R., and Shi, X. In Characterization of Dendrimer–Gold Nanocomposite<br />

Materials, Materials Research Society Symposium Proceedings, Warrendale, PA, 2005; Sanchez,<br />

C., Schubert, U., Laine, R. M., Chujo, Y., Materials Research Society, Warrendale, PA, EE13.33,<br />

pp. 1–6, 2005.<br />

66. Esumi, K., Dendrimers for nanoparticle synthesis and dispersion stabilization, Top. Curr. Chem.,<br />

227, 31–52, 2003.<br />

67. Seo, Y.-S., Kim, K. S., Shin, K., White, H., Rafailovich, M., Sokolov, J., Lin, B., Kim, H. J., Zhang,<br />

C., and Balogh, L., Morphology of amphiphilic gold/dendrimer nanocomposite monolayers, Langmuir,<br />

18 (15), 5927–5932, 2002.<br />

68. Tomalia, D. A., Baker, H., Dewald, J., Hall, M., and Kallos, G., Dendritic macromolecules:<br />

Synthesis of starburst dendrimers, Macromolecules, 19, 2466–2468, 1986.<br />

69. Uppuluri, S., Swanson, D. R., Piehler, L. T., Li, J., Hagnauer, G. L., and Tomalia, D. A., Core-shell<br />

tecto(dendrimers): I. Synthesis and characterization of saturated shell models, Adv. Mater., 12 (11),<br />

796–800, 2000.<br />

70. Li, J., Swanson, D. R., Qin, D., Brothers, H. M., Piehler, L. T., Tomalia, D., and Meier, D. J.,<br />

Characterizations of core-shell tecto-(dendrimer) molecules <strong>by</strong> tapping mode atomic force<br />

microscopy, Langmuir, 15 (21), 7347–7350, 1999.<br />

71. Tomalia, D. A., Piehler, L. T., Durst, H. D., and Swanson, D. R., Partial shell-filled core-shell<br />

tecto(dendrimers): A strategy to surface differentiated nano-clefts and cusps, Proc. Natl Acad.<br />

Sci. U.S.A., 99 (8), 5081–5087, 2002.<br />

72. Shi, X., Patri, A. K., Lesniak, W., Islam, M. T., Zhang, C., Baker, J. R., and Balogh, L. P., Analysis of<br />

poly(amidoamine)-succinamic acid dendrimers <strong>by</strong> slab-gel electrophoresis and capillary zone<br />

electrophoresis, Electrophoresis, 26, 2960–2967, 2005.<br />

73. Islam, M. T., Shi, X., Balogh, L. P., and Baker, J. R., HPLC separation of different generations of<br />

poly(amidoamine) dendrimers modified with various terminal groups, Anal. Chem., 77, 2063–2070,<br />

2005.<br />

74. Shi, X., Lesniak, W., Islam, M. T., Muñiz, M. C., Balogh, L. P., and Baker, J. R., Comprehensive<br />

characterization of surface-functionalized poly(amidoamine) dendrimers with acetamide, hydroxyl,<br />

and carboxyl groups, Colloids Surf., A: Physicochem. Eng. Aspects, 272, 139–150, <strong>2006</strong>.<br />

75. Shi, X., Bányai, I., Islam, M. T., Lesniak, W., Davis, D. Z., Baker, J. R., and Balogh, L. P.,<br />

Generational, skeletal and substitutional diversities in generation one poly(amidoamine) dendrimers,<br />

Polymer, 46 (9), 3022–3034, 2005.<br />

76. Lesniak, W., Bielinska, A. U., Sun, K., Janczak, K. W., Shi, X., Baker, J. R., and Balogh, L. P.,<br />

Silver/dendrimer nanocomposites as biomarkers: Fabrication, characterization, in vitro toxicity, and<br />

intracellular detection, Nano Lett., 5 (11), 2123–2130, 2005.<br />

77. Nigavekar, S. S., Sung, L. Y., Llanes, M., El-Jawahri, A., Lawrence, T. S., Becker, C. W., Balogh, L.,<br />

and Khan, M. K., 3H dendrimer nanoparticle organ/tumor distribution, Pharm. Res., 21 (3), 476–483,<br />

2004.<br />

78. Malik, N., Wiwattanapatapee, R., Klopsch, R., Lorenz, K., Frey, H., Weener, J. W., Meijer, E. W.,<br />

Paulus, W., and Duncan, R., Dendrimers: Relationship between structure and biocompatibility<br />

in vitro, and preliminary studies on the biodistribution of 125I-labelled polyamidoamine dendrimers<br />

in vivo, J. Control. Release, 65 (1–2), 133–148, 2000.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


590<br />

Nanotechnology for Cancer Therapy<br />

79. Kukowska-Latallo, J. F., Candido, K. A., Cao, Z., Nigavekar, S. S., Majoros, I. J., Thomas, T. P.,<br />

Balogh, L. P., Khan, M. K., and Baker, J. R. Jr., Nanoparticle targeting of anticancer drug improves<br />

therapeutic response in animal model of human epithelial cancer, Cancer Res., 65 (12), 5317–5324,<br />

2005.<br />

80. Tang, M. X., Redemann, C. T., and Szoka, F. C., In vitro gene delivery <strong>by</strong> degraded polyamidoamine<br />

dendrimers, Bioconjug. Chem., 7, 703–714, 1996.<br />

81. Rosi, N. L. and Mirkin, C. A., Nanostructures in biodiagnostics, Chem. Rev., 105, 1547–1562, 2005.<br />

82. Burda, C., Chen, X., Narayanan, R., and El-Sayed, M. A., Chemistry and properties of nanocrystals<br />

of different shapes, Chem. Rev., 105, 1025–1102, 2005.<br />

83. Bonkhoff, H. and Remberger, K., Widespread distribution of nuclear androgen receptors in the basal<br />

cell layer of the normal and hyperplastic human prostate, Virchows Arch. A Pathol. Anat. Histopathol.,<br />

422 (1), 35–38, 1993.<br />

84. Pasqualini, R. and Ruoslahti, E., Organ targeting in vivo using phage display peptide libraries,<br />

Nature, 380 (6572), 364–366, 1996.<br />

85. Rajotte, D. and Ruoslahti, E., Membrane dipeptidase is the receptor for a lung-targeting peptide<br />

identified <strong>by</strong> in vivo phage display, J. Biol. Chem., 274 (17), 11593–11598, 1999.<br />

86. Pasqualini, R., Koivunen, E., Kain, R., Lahdenranta, J., Sakamoto, M., Stryhn, A., Ashmun, R. A.,<br />

Shapiro, L. H., Arap, W., and Ruoslahti, E., Aminopeptidase N is a receptor for tumor-homing<br />

peptides and a target for inhibiting angiogenesis, Cancer Res., 60 (3), 722–727, 2000.<br />

87. Arap, W., Pasqualini, R., and Ruoslahti, E., Cancer treatment <strong>by</strong> targeted drug delivery to tumor<br />

vasculature in a mouse model, Science, 279 (5349), 377–380, 1998.<br />

88. Pasqualini, R., Koivunen, E., and Ruoslahti, E., Alpha v integrins as receptors for tumor targeting <strong>by</strong><br />

circulating ligands, Nat. Biotechnol., 15 (6), 542–546, 1997.<br />

89. Burg, M. A., Pasqualini, R., Arap, W., Ruoslahti, E., and Stallcup, W. B., NG2 proteoglycan-binding<br />

peptides target tumor neovasculature, Cancer Res., 59 (12), 2869–2874, 1999.<br />

90. Eller<strong>by</strong>, H. M., Arap, W., Eller<strong>by</strong>, L. M., Kain, R., Andrusiak, R., Rio, G. D., and Krajewski, S.,<br />

Anti-cancer activity of targeted pro-apoptotic peptides, Nat. Med., 5 (9), 1032–1038, 1999. http://<br />

www.java/Propub/medicine/nm0999_1032.fulltextjava/Propub/medicine/nm0999_1032.abstract.<br />

91. Dmitriev, I., Krasnykh, V., Miller, C. R., Wang, M., Kashentseva, E., Mikheeva, G., Belousova, N.,<br />

and Curiel, D. T., An adenovirus vector with genetically modified fibers demonstrates expanded<br />

tropism via utilization of a coxsackievirus and adenovirus receptor-independent cell entry<br />

mechanism, J. Virol., 72 (12), 9706–9713, 1998.<br />

92. DeNardo, S. J., Burke, P. A., Leigh, B. R., O’Donnell, R. T., Miers, L. A., Kroger, L. A., and<br />

Goodman, S. L., Neovascular targeting with cyclic RGD peptide (cRGDf-ACHA) to enhance<br />

delivery of radioimmunotherapy, Cancer Biother. Radiopharm., 15 (1), 71–79, 2000.<br />

93. Janssen, M. L., Oyen, W. J., Dijkgraaf, I., Massuger, L. F., Frielink, C., Edwards, D. S., Rajopadhye,<br />

M., Boonstra, H., Corstens, F. H., and Boerman, O. C., Tumor targeting with radiolabeled alpha(v)beta(3)<br />

integrin binding peptides in a nude mouse model, Cancer Res., 62 (21), 6146–6151, 2002.<br />

94. Ogawa, M., Hatano, K., Oishi, S., Kawasumi, Y., Fujii, N., Kawaguchi, M., Doi, R., Imamura, M.,<br />

Yamamoto, M., Ajito, K. et al., Direct electrophilic radiofluorination of a cyclic RGD peptide for<br />

in vivo alpha(v)beta(3) integrin related tumor imaging(1), Nucl. Med. Biol., 30 (1), 1–9, 2003.<br />

95. Haubner, R., Wester, H. J., Weber, W. A., Mang, C., Ziegler, S. I., Goodman, S. L., Senekowitsch-<br />

Schmidtke, R., Kessler, H., and Schwaiger, M., Noninvasive imaging of alpha(v)beta3 integrin<br />

expression using 18F-labeled RGD-containing glycopeptide and positron emission tomography,<br />

Cancer Res., 61 (5), 1781–1785, 2001.<br />

96. Sivolapenko, G. B., Skarlos, D., Pectasides, D., Stathopoulou, E., Milonakis, A., Sirmalis, G.,<br />

Stuttle, A., Courtenay-Luck, N. S., Konstantinides, K., and Epenetos, A. A., Imaging of metastatic<br />

melanoma utilising a technetium-99m labelled RGD-containing synthetic peptide, Eur. J. Nucl.<br />

Med., 25 (10), 1383–1389, 1998.<br />

97. Liu, S., Edwards, D. S., Ziegler, M. C., Harris, A. R., Hemingway, S. J., and Barrett, J. A., 99mTclabeling<br />

of a hydrazinonicotinamide-conjugated vitronectin receptor antagonist useful for imaging<br />

tumors, Bioconjug. Chem., 12 (4), 624–629, 2001.<br />

98. Folkman, J., Tumor Angiogenesis, In Cancer Medicine, Holland, J. F., Bast, R. C., Morton, D. L.,<br />

Frei, E., Kufe, D. W., and Weichselbaum, R. R., Eds., Vol. 1, Williams and Wilkens, Baltimore, MD,<br />

pp. 181–204, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Dendrimer Nanocomposites for Cancer Therapy 591<br />

99. Jain, R. K., Determinants of tumor blood flow: A review, Cancer Res., 48 (10), 2641–2658, 1998.<br />

100. Ribatti, D., Vacca, A., and Dammacco, F., The role of the vascular phase in solid tumor growth: A<br />

historical review, Neoplasia, 1 (4), 293–302, 1999.<br />

101. Schlingemann, R. O., Rietveld, F. J., Kwaspen, F., van de Kerkhof, P. C., de Waal, R. M., and Ruiter,<br />

D. J., Differential expression of markers for endothelial cells, pericytes, and basal lamina in the<br />

microvasculature of tumors and granulation tissue, Am. J. Pathol., 138 (6), 1335–1347, 1991.<br />

102. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent<br />

smancs, Cancer Res., 46 (12 Pt 1), 6387–6392, 1986.<br />

103. Duncan, R., Malik, N., Richardson, S., and Ferruti, P., Polymer conjugates for anti-cancer agent and<br />

DNA delivery, Polym. Prepr. (Am. Chem. Soc. Div., Polym. Chem.), 39, 180, 1998.<br />

104. Muggia, F. M., Doxorubicin–polymer conjugates: Further demonstration of the concept of enhanced<br />

permeability and retention, Clin. Cancer Res., 5 (1), 7–8, 1999.<br />

105. Hainfeld, J. F., Foley, C. J., Srivastava, S. C., Mausner, L. F., Feng, N. I., Meinken, G. E., and<br />

Steplewski, Z., Radioactive gold cluster immunoconjugates: Potential agents for cancer therapy, Int.<br />

J. Rad. Appl. Instrum. B, 17 (3), 287–294, 1990.<br />

106. Sharma, V., Luker, G. D., and Piwnica-Worms, D., Molecular imaging of gene expression and<br />

protein function in vivo with PET and SPECT, J. Magn. Reson. Imaging, 16 (4), 336–351, 2002.<br />

107. Berning, D. E., Katti, K. V., Volkert, W. A., Higginbotham, C. J., and Ketring, A. R., 198Au-labeled<br />

hydroxymethyl phosphines as models for potential therapeutic pharmaceuticals, Nucl. Med. Biol.,25<br />

(6), 577–583, 1998.<br />

108. Brechbiel, M. W., Gansow, O. A., Wu, C., Garmestani, K., Yordanov, A., Deal, K., and Chappell, L.,<br />

In Chelated Metal Ions for Therapeutic and Diagnostic Applications, 215th ACS Natioinal Meeting,<br />

Dallas, March 29–April 2, ACS, Washington, DC, Dallas, NUCL-106, 1998.<br />

109. Wiener, E. C., Brechbiel, M. W., Brothers, H., Magin, R. L., Gansow, O. A., Tomalia, D. A., and<br />

Lauterbur, P. C., Dendrimer-based metal chelates: A new class of magnetic resonance imaging<br />

contrast agents, Magn. Reson. Med., 31 (1), 1–8, 1994.<br />

110. Wu, C., Brechbiel, M. W., Kozak, R. W., and Gansow, O. A., Metal–chelate–dendrimer–antibody<br />

constructs for use in radioimmunotherapy and imaging, Bioorg. Med. Chem. Lett., 4 (3), 449–454,<br />

1994.<br />

111. Wiener, E. C., Konda, S., Shadron, A., Brechbiel, M., and Gansow, O., Targeting dendrimer-chelates<br />

to tumors and tumor cells expressing the high-affinity folate receptor, Investig. Radiol., 32 (12),<br />

748–754, 1997.<br />

112. Brechbeil, M., J. Nucl. Med., 35 (5 suppl. S), 62, 1994.<br />

113. Barth, R. F., Soloway, A. H., and Brugger, R. M., Boron neutron capture therapy of brain tumors:<br />

Past history, current status, and future potential, Cancer Investig., 14 (6), 534–550, 1996.<br />

114. Margerum, L. D., Campion, B. K., Koo, M., Shargill, N., Lai, J., Marumoto, A., and Sontum, P. C.,<br />

J. Alloys Compd., 249 (1–2), 185–190, 1997.<br />

115. Von Kleist, S., Ten years of tumor imaging with labelled antibodies, In Vivo, 7 (6B), 581–584, 1993.<br />

116. Blankenberg, F. G. and Strauss, H. W., Nuclear medicine applications in molecular imaging,<br />

J. Magn. Reson. Imaging, 16 (4), 352–361, 2002.<br />

117. Kwekkeboom, D., Krenning, E. P., and de Jong, M., Peptide receptor imaging and therapy, J. Nucl.<br />

Med., 41 (10), 1704–1713, 2000.<br />

118. Ang, E. S. and Sundram, F. X., Prospects for tumour imaging with radiolabelled antibodies, Ann.<br />

Acad. Med. Singapore, 22 (5), 776–784, 1993.<br />

119. Deshpande, S. V., DeNardo, S. J., Kukis, D. L., Moi, M. K., McCall, M. J., DeNardo, G. L., and<br />

Meares, C. F., Yttrium-90-labeled monoclonal antibody for therapy: Labeling <strong>by</strong> a new macrocyclic<br />

bifunctional chelating agent, J. Nucl. Med., 31 (4), 473–479, 1990.<br />

120. Beaumier, P. L., Venkatesan, P., Vanderheyden, J. L., Burgua, W. D., Kunz, L. L., Fritzberg, A. R.,<br />

Abrams, P. G., and Morgan, A. C. Jr., 186Re radioimmunotherapy of small cell lung carcinoma<br />

xenografts in nude mice, Cancer Res., 51 (2), 676–681, 1991.<br />

121. Chalandon, Y., Mach, J. P., Pelegrin, A., Folli, S., and Buchegger, F., Combined radioimmunotherapy<br />

and chemotherapy of human colon carcinoma grafted in nude mice, advantages and<br />

limitations, Anticancer Res., 12 (4), 1131–1139, 1992.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


592<br />

Nanotechnology for Cancer Therapy<br />

122. Cheung, N. K., Landmeier, B., Neely, J., Nelson, A. D., Abramowsky, C., Ellery, S., Adams, R. B.,<br />

and Miraldi, F., Complete tumor ablation with iodine 131-radiolabeled disialoganglioside GD2specific<br />

monoclonal antibody against human neuroblastoma xenografted in nude mice, J. Natl<br />

Cancer Inst., 77 (3), 739–745, 1986.<br />

123. Esteban, J. M., Schlom, J., Mornex, F., and Colcher, D., Radioimmunotherapy of athymic mice<br />

bearing human colon carcinomas with monoclonal antibody B72.3: Histological and autoradiographic<br />

study of effects on tumors and normal organs, Eur. J. Cancer Clin. Oncol., 23 (6),<br />

643–655, 1987.<br />

124. Balogh,, L. P., Nigavekar, S. S., Cook, A. C., Minc, L., and Khan, M. K., Development of<br />

dendrimer–gold radioactive nanocomposites to treat cancer microvasculature, PharmaChem, 2<br />

(4), 94–99, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


29<br />

CONTENTS<br />

Applications of Liposomal<br />

Drug Delivery Systems<br />

to Cancer Therapy<br />

Alberto A. Gabizon<br />

29.1 Introduction ....................................................................................................................... 595<br />

29.2 The Challenge of Cancer Therapy ................................................................................... 596<br />

29.3 The Rationale for the Use of Liposomal Drug Carriers .................................................. 597<br />

29.4 Differential Factors on the Pharmacokinetics of Low-Molecular-<br />

Weight Drugs and Liposomes .......................................................................................... 597<br />

29.5 Correlation Between Liposome Formulation and Liposome<br />

Pharmacokinetics-Stealth Liposomes ............................................................................... 598<br />

29.6 Preclinical Observations with Liposomal Anthracyclines ............................................... 600<br />

29.7 Liposomal Anthracyclines in the Clinic—Doxil (PLD) .................................................. 602<br />

29.8 Development of Other Liposome-Entrapped Cytotoxic Agents...................................... 605<br />

29.9 Future Avenues of Liposome Development—Receptor<br />

Targeted Liposomes.......................................................................................................... 606<br />

29.10 Conclusions ....................................................................................................................... 607<br />

References.................................................................................................................................... 607<br />

29.1 INTRODUCTION<br />

In the last two decades, major advances in the use of injectable drug delivery systems for the<br />

treatment of cancer have occurred. These include macromolecular conjugates, liposomes, and other<br />

nanoparticles. Poly(ethylene glycol) (PEG)-modified liposomal doxorubicin (Doxil, Caelyx) was<br />

the first liposomal anti-cancer drug to be approved <strong>by</strong> the Food and Drug Administration, 1 whereas<br />

paclitaxel albumin-bound particle suspension (ABI007, Abraxane) was recently approved for the<br />

treatment of metastatic breast cancer. 2 The ultimate goal behind the development of these sophisticated<br />

drug delivery systems is to improve the efficacy and decrease the side effects of new and old<br />

anti-cancer drugs. To achieve this goal, changes in the pharmacokinetics, biodistribution, and<br />

bioavailability profile leading to a positive impact on the drug pharmacodynamics are required.<br />

This chapter will focus on injectable liposome-based drug delivery systems of anti-cancer<br />

drugs, the most widely used drug nanoparticle in cancer. Liposomes are vesicles with an<br />

aqueous interior surrounded <strong>by</strong> one or more concentric bilayers of phospholipids with a diameter<br />

ranging from a minimal diameter of w30 nm to several microns. 3 However, for injectable<br />

clinical applications, practically all liposome formulations are in the submicron ultrafilterable<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

595


596<br />

range (!200 nm size) and can be considered as nanosize particulate systems. Liposomes are<br />

spontaneously formed when amphiphilic lipids such as phospholipids are dispersed in water.<br />

The ensuing structures are physically stable supramolecular assemblies, and, unlike polymerized<br />

particles, they are not covalently bound. Whether the drug is encapsulated in the aqueous core or in<br />

the surrounding bilayer of the liposome is dependent on the characteristics of the drug and the<br />

encapsulation process. 4 In general, water-soluble drugs are encapsulated within the central aqueous<br />

core, whereas lipid-soluble drugs are incorporated directly into the lipid membrane.<br />

The current trend is to classify liposomes into a class of pharmaceutical devices in the nanoscale<br />

range engineered <strong>by</strong> physical and/or chemical means, and referred to as nanomedicines. 5 The<br />

field of nanomedicines is rapidly evolving and aims at increased sophistication in the design of<br />

nanosize devices and their interactions with cellular targets at the nanoscale level. In fact, liposomes<br />

are the first generation of nanomedicines approved for treatment of cancer (Doxil, described<br />

later in this chapter) and fungal infections (Ambisome, containing amphotericin B). Current liposome<br />

formulations involve a slow drug release system and a passive targeting process known as<br />

enhanced permeability and retention (EPR) that will be discussed later in this chapter.<br />

29.2 THE CHALLENGE OF CANCER THERAPY<br />

Nanotechnology for Cancer Therapy<br />

Current understanding of the molecular processes underlying the pathologic behavior of cancer<br />

cells has progressed enormously in the last decade. 6 Of particular relevance to cancer targeting is<br />

the fact that a number of receptors, mostly growth factor receptors, have been found to be overexpressed<br />

in tumor cells and to play an important role as catalysts of growth. Receptor profiling of<br />

tumors may offer a potential Achilles heel for targeting specific ligands or antibodies with or<br />

without delivery of a cytotoxic drug cargo. 7 In addition, the pathophysiology of tumor neovasculature<br />

and the interaction of tumor with stroma have been recognized as processes that play a major<br />

role in tumor development. Cancer is ultimately a disease caused <strong>by</strong> somatic gene mutations that<br />

result in the transformation of a normal cell into a malignant tumor cell. Eventually, the tumor cell<br />

phenotype progresses along three major steps: 8<br />

1. Increased proliferation rate and/or decreased apoptosis, causing an increase of tumor<br />

cell mass.<br />

2. Invasion of surrounding tissues and switch on of angiogenesis. This is a critical step that<br />

differentiates in situ, non-invasive tumors with no metastatic potential from invasive<br />

tumors with metastatic and life-threatening potential.<br />

3. Metastases, i.e., abnormal migration of tumor cells from the primary tumor site via blood<br />

vessels or lymphatics to distant organs with formation of secondary tumors. Most<br />

commonly, this is the process that causes death of the host because of disruption of<br />

the function of vital organs or systems (brain, lung, liver, kidney, bone marrow, coagulation,<br />

intestinal passage, and other).<br />

Despite formidable advances in clinical imaging, the diagnosis of a tumor mass requires usually<br />

the presence of a nodule of w10 mm diameter representing a cluster of 10 9 cells.* Because the<br />

lethal tumor burden is in the order of 10 12 cells in most cancer patients, this implies that tumors<br />

have already gone through 3/4 of their doubling cell expansion process <strong>by</strong> the time of clinical<br />

diagnosis. As a result, significant heterogeneity and phenotypic diversity are already present in<br />

* This does not apply to superficial skin tumors that can be recognized often when they are only 2–3 mm diameter and<br />

contain clusters of w10 7 cells. Occasionally, modern imaging techniques (high resolution CT scan, MRI) can detect deepseated<br />

lesions with suspected cancer features that are smaller than 10 mm.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 597<br />

most diagnosed cancers, posing a major therapeutic challenge as a result of the development of<br />

metastatic ability and drug resistance.<br />

In addition to the lack of specificity of chemotherapeutic (cytotoxic) agents, a number of<br />

physiologic factors can seriously limit the efficiency of drug distribution from plasma to tumors<br />

and neutralize their effects. These include competition for drug uptake of well-perfused tissues such<br />

as liver and kidneys, rapid glomerular filtration and urinary excretion of low-molecular-weight<br />

drugs, protein binding with drug inactivation (e.g., cisplatin), and stability problems in biological<br />

fluids (e.g., hydrolysis of nitrosoureas, opening of lactone ring of camptothecin analogs).<br />

29.3 THE RATIONALE FOR THE USE OF LIPOSOMAL DRUG CARRIERS<br />

The rationale for the use of liposomes in cancer drug delivery is based on the following pharmacological<br />

principles: 1<br />

1. Slow drug release: drug bioavailability depends on drug release from liposomes. Entrapment<br />

of drug in liposomes will slow down drug release and reduce renal clearance to a<br />

variable extent. Slow release may range from a mere blunting of the peak plasma levels<br />

of free drug to a sustained release of drug, mimicking continuous infusion. These pharmacokinetic<br />

changes may have important pharmacodynamic consequences with regard<br />

to toxicity and efficacy of the liposome delivered agents.<br />

2. Site avoidance of specific tissues: the biodistribution pattern of liposomes may lead to a<br />

relative reduction of drug concentration in tissues specifically sensitive to the delivered<br />

drug. This may have implications with regard to the therapeutic window of various<br />

cytotoxic drugs such as the cardiotoxic anthracyclines, provided that anti-tumor efficacy<br />

is not negatively affected.<br />

3. Accumulation in tumors: prolongation of the circulation time of liposomes results in<br />

significant accumulation in tissues with increased vascular permeability. This is often the<br />

case of tumors, especially in those areas with active neoangiogenesis. 9 Tumor localization<br />

of long-circulating liposomes such as pegylated liposomes, sometimes referred to<br />

as Stealth or sterically-stabilized, 10 is a passive targeting effect, enabling substantial<br />

accumulation of liposome-encapsulated drug in the interstitial fluid at the tumor site, 11<br />

a phenomenon sometimes referred as EPR effect.<br />

29.4 DIFFERENTIAL FACTORS ON THE PHARMACOKINETICS OF LOW-<br />

MOLECULAR-WEIGHT DRUGS AND LIPOSOMES<br />

There are a number of differential effects of physiologic factors on clearance and biodistribution of<br />

low-molecular-weight drugs and nanoparticles such as liposomes.<br />

† Protein binding: low-molecular-weight drugs may be inactivated and/or irreversibly<br />

bound <strong>by</strong> plasma proteins, reducing the bioavailability toward cellular target molecules.<br />

This is the case of cisplatin, a widely used anti-cancer cytotoxic drug. In the case of<br />

nanoparticles, plasma proteins can adsorb to their surface, a process known as opsonization<br />

that results in tagging the particle for recognition and removal <strong>by</strong> macrophages. In<br />

addition, protein binding to the liposome surface may de-stabilize the bilayer and accelerate<br />

the leakage of liposome contents. PEG coating (PEGylation) of liposomes reduces<br />

opsonization and the effects associated with it.<br />

† Reticulo-endothelial system (RES) clearance: it is unimportant for low-molecularweight<br />

drugs, but it plays a major role in the clearance of nanoparticles, reducing the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


598<br />

Nanotechnology for Cancer Therapy<br />

TABLE 29.1<br />

Parameters Affecting Delivery of Liposomal Drugs to Tumors<br />

Tumor Factors Liposome Factors<br />

Blood flow Long circulation time<br />

Vascular permeability Stability (drug retention)<br />

Interstitial pressure Small vesicle size<br />

Phagocytic activity<br />

For a detailed discussion of these factors, see text.<br />

Saturation of the RES<br />

fraction available for distribution to tumor tissue. Kupffer cell macrophages lining the<br />

liver sinusoids remove opsonized liposomes and other nanoparticles from circulation and<br />

represent a major factor in the clearance of particulate carriers.<br />

† Glomerular filtration: unless they become protein-bound, low-molecular-weight drugs<br />

can be filtered out <strong>by</strong> kidney glomeruli. In contrast, liposomes and other nanoparticles<br />

are non-filterable because their radius exceeds the glomerular filterable threshold size.<br />

† Microvascular permeability: enhanced microvascular permeability with fenestrations in<br />

capillaries and post-capillary venules is critical for extravasation of nanoparticles from<br />

the blood stream to the interstitial fluid of target tissues. The presence of fenestrations is<br />

irrelevant for tissue delivery of small molecules.<br />

† Extravascular transport: diffusion is the predominant mechanism of transport for small<br />

molecules. In contrast, convective transport plays a major role in the extravascular<br />

movement of nanoparticles for which diffusion rates are very slow. 12 Large tumors<br />

tend to develop high interstitial pressure that reduces the rate of convective transport.<br />

In fact, in an animal model, it has been shown that liposomes accumulate significantly<br />

less in larger tumors on a per gram tissue basis. 13 In agreement with this, large tumor size<br />

predicts poor response to liposome-delivered chemotherapy in ovarian cancer. 14<br />

Table 29.1 lists a number of tumor and liposome factors that play an important role in the<br />

delivery of liposomal drugs. On the tumor side, a rich blood flow and highly permeable microvascular<br />

bed will increase the probability of liposome deposition, whereas a high interstitial fluid<br />

pressure is likely to reduce the movement of molecules and particles into the tumor compartment.<br />

On the liposome side, avoiding drug leakage and prolonging the circulation time will result in more<br />

liposomes reaching the tumor vascular bed with an intact drug payload, and a small vesicle size will<br />

facilitate extravasation through the endothelium gaps or fenestrations. There are also data indicating<br />

that saturation of the RES will prolong circulation time and indirectly enhance liposome<br />

deposition in tumors. 15<br />

29.5 CORRELATION BETWEEN LIPOSOME FORMULATION AND LIPOSOME<br />

PHARMACOKINETICS-STEALTH LIPOSOMES<br />

In 1971, Gregoriadis et al. 16 published the first research work where liposomes were used as drug<br />

carriers for medical applications. This initial study led to growing interest in liposomes, and many<br />

laboratories began examining liposome pharmacokinetics and biodistribution in animals as well as<br />

in vitro stability in serum. The early liposome work was mostly based on formulations composed of<br />

neutral egg lecithin (PC), often in combination with negatively or positively charged lipids. These<br />

liposomes were found to rapidly release a large fraction their encapsulated contents in circulation<br />

and were quickly removed from circulation <strong>by</strong> macrophages of the RES residing in liver and spleen.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 599<br />

Reformulation with high phase-transition temperature (Tm) lipids (distearoyl-PC, dipalmitoyl-PC,<br />

sphingomyelin) and an addition of cholesterol led to improved retention of liposome contents<br />

and prolongation of circulation time, especially when the vesicles were properly downsized to<br />

!100 nm diameter. However, these relatively improved liposome formulations would still largely<br />

accumulate in the RES, and a greater improvement in circulation half-life appeared to be required<br />

for cancer targeting. Surface modifications of liposomes that could reduce the RES affinity were<br />

investigated based on the erythyrocyte paradigm where<strong>by</strong> a layer of carbohydrate groups prolong<br />

circulation for nearly three months. A number of glycolipids such as monosialoganglioside (GM1),<br />

phosphatidyl-inositol, and cerebroside sulfate were included in the formulations and, indeed,<br />

extended liposome circulation time. 17,18 However, a major leap forward took place when the<br />

hydrophilic polymer PEG that was known to reduce immunogenicity and prolong circulation<br />

time when attached to enzymes and growth factors was introduced into liposomes in the early<br />

1990s. PEG that is inexpensive as a result of easy synthesis and that could be prepared in high purity<br />

and large quantities had distinct advantages over the other glycolipid surface modifiers. Addition of<br />

a conjugate of PEG with a lipid anchor, distearoyl-phosphatidylethanolamine (PEG–DSPE), to the<br />

liposomal formulation was shown to significantly prolong liposome circulation time 10,19–22 and<br />

formed a pivotal element of the pharmaceutical development of the DOXIL formulation described<br />

thereafter in this chapter. Because of their ability to avoid/escape RES clearance mechanisms, PEGcoated<br />

liposomes have been coined Stealth* liposomes. 23 In general, the pharmacokinetic variables<br />

of the liposomes depend on the liposomes’ physicochemical characteristics such as size, surface<br />

charge, membrane lipid packing, steric stabilization as conferred <strong>by</strong> PEG coating, dose, and route<br />

of administration.<br />

It should be noted that the pharmacokinetic disposition of liposome-associated agents is dependent<br />

on the carrier until the drug is released from the carrier. Therefore, the pharmacology and<br />

pharmacokinetics of drugs delivered in liposomes are complex, and ideally, studies must be done to<br />

evaluate the disposition of the entrapped form of the drug and of the released active drug. A<br />

prerequisite for these types of studies is to develop the proper methodology to discriminate<br />

between both forms of the drug in plasma.<br />

In parallel to the development of stable formulations with long circulation half-life, it was soon<br />

realized that a prolonged residence time in circulation was a critical pharmacokinetic factor for<br />

liposome deposition in tumors and that there was a strong correlation between liposome circulation<br />

time and tumor uptake. 18 A number of studies have addressed the mechanism of liposome accumulation<br />

in tumors. Microscopic observations with colloidal gold-labeled liposomes 24 and<br />

morphologic studies with fluorescent liposomes in the skin-fold chamber model 25 have demonstrated<br />

that liposomes extravasate into the tumor extracellular fluid through gaps in tumor<br />

microvessels and are found predominantly in the perivascular area with minimal uptake <strong>by</strong><br />

tumor cells. Studies with ascitic tumors demonstrate a steady extravasation process of long circulating<br />

liposomes into the ascitic fluid with gradual release of drug followed <strong>by</strong> drug diffusion into<br />

the ascitic cellular compartment. 26,27 The process underlying the preferential tumor accumulation<br />

of liposomes as well as other macromolecular and particulate carriers is known as EPR effect. 28<br />

This is a passive and non-specific process, resulting from increased microvascular permeability and<br />

defective lymphatic drainage in tumors creating an in situ depot of liposomes in the tumor interstitial<br />

fluid. Circulating liposomes cross the leaky tumor vasculature, moving from plasma into the<br />

interstitial fluid of tumor tissue following convective transport and diffusion processes. Although<br />

convective transport of plasma fluid occurs also in normal tissues, the continuous, non-fenestrated<br />

endothelium and basement membrane prevents the extravasation of liposomes. EPR is relatively a<br />

slow process where long-circulating liposomes possess a distinct advantage because of the repeated<br />

* Stealth is a registered trademark of Alza Corp., Mountain View, CA.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


600<br />

Nanotechnology for Cancer Therapy<br />

passage through the tumor microvascular bed and their high concentration in plasma during an<br />

extended period of time.<br />

For any intra-vascular drug carrier device to access the tumor cell compartment and interact<br />

with tumor cell receptors, it must first cross the vascular endothelium and diffuse into the interstitial<br />

fluid because, with few exceptions, tumor cells and their surface receptors are not directly exposed<br />

to the blood stream. Therefore, the EPR effect is not only important for the tumor accumulation of<br />

non-targeted liposomes but also for that of ligand-targeted liposomes. This has led to the assertion<br />

that the extravasation process is the rate-limiting step of liposome accumulation in tumors. 29<br />

Experimental data with targeted and non-targeted liposomes have so far lent consistent support<br />

to this hypothesis. 30<br />

In most instances, delivery of drug to tumor cells depends on the release of drug from liposomes<br />

in the interstitial fluid because liposomes are seldom taken up <strong>by</strong> tumor cells unless they are tagged<br />

with specific ligands. The factors controlling this process and its kinetics are not well understood<br />

and may vary among tissues, depending on the liposome formulation in question. In the case of<br />

remote-loaded formulations, that is often the case of anthracyclines, a gradual loss of the liposome<br />

gradient retaining the drug in addition to disruption of the integrity of the liposome bilayer <strong>by</strong><br />

phospholipases may be involved in the release process. Uptake <strong>by</strong> tumor-infiltrating macrophages<br />

could also contribute to liposomal drug release. In any case, once the drug is released from<br />

liposomes, it will freely diffuse through the interstitial tumor space and reach deep layers of<br />

tumor cells. This is an inherent advantage of this delivery system as opposed to covalently<br />

bound drug-carrier systems. It is also a critical factor for the success of the liposomal drug approach<br />

because most of the liposomes appear to remain in interstitial spaces immediately surrounding the<br />

blood vessels, 24,25 and therefore would not be able to interact with more than one layer of tumor<br />

cells. As a result, the ability of the liposome to effectively carry the anti-cancer agent to the tumor<br />

and the ability to release the drug at a satisfactory rate in the tumor extracellular fluid are equally<br />

important factors in determining the anti-tumor effect of liposomal-encapsulated anticancer<br />

agents.<br />

The EPR effect has been confirmed in a variety of implanted tumor models. Its validity<br />

regarding human tumors and, particularly, cancer metastases is unclear as yet. One point of<br />

concern is the fact that interstitial fluid pressure increases in most tumors once they grow above<br />

a certain size threshold, 31 therefore hindering extravascular transport and liposome delivery. Unfortunately,<br />

there is a paucity of imaging studies in cancer patients with radiolabeled liposomes. One<br />

of the few studies with radiolabeled pegylated liposomes demonstrated significant liposome<br />

accumulation based on tumor imaging findings in 15/17 patients tested. 32 In another study where<br />

tumor metastases and normal muscle tissues of two breast cancer patients were examined for<br />

doxorubicin concentration after injection of pegylated liposomal doxorubicin (PLD), liposomal<br />

drug was found at 10-fold greater concentration in tumor as compared to muscle. 33 Another<br />

important piece of work in this area is the study of Northfelt et al. 34 that pointed to an enhanced<br />

deposition of drug in Kaposi’s sarcoma (KS) skin lesions of patients receiving PLD as compared to<br />

normal skin of the same patients and to doxorubicin concentration in KS biopsies of patients<br />

receiving free doxorubicin.<br />

29.6 PRECLINICAL OBSERVATIONS WITH LIPOSOMAL ANTHRACYCLINES<br />

The drugs most frequently incorporated and evaluated in liposomal formulations are anthracyclines,<br />

including doxorubicin and daunorubicin. The choice of doxorubicin <strong>by</strong> many of the early<br />

research groups examining the role of liposomes as drug carriers in cancer chemotherapy stems<br />

from its broad spectrum of anti-tumor activity on the one hand, and its disturbing cumulative doselimiting<br />

cardiac toxicity. Anthracyclines such as doxorubicin and daunorubicin cause acute toxic<br />

side effects, including bone marrow depression, alopecia, and stomatitis, and they are dose limited<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 601<br />

<strong>by</strong> a serious and mostly irreversible characteristic cardiomyopathy. 35 The first study describing the<br />

encapsulation of anthracyclines into liposomes appeared in 1979. 36 Work from various research<br />

groups followed soon after, supporting the general principle that liposomal formulations reduced<br />

the toxicity of anthracyclines in animal models. 37<br />

Based on the Stealth concept and using an elegant drug-loading mechanism, a formulation of<br />

PLD known as DOXIL (Caelyx in Europe) has been developed (Figure 29.1). The loading<br />

mechanism, coined remote (active) loading, is based on an ammonium sulfate gradient, leading<br />

to marked accumulation of doxorubicin inside the aqueous phase of the liposome (w15,000<br />

doxorubicin molecules/vesicle). At high concentration, the entrapped drug forms a crystallinelike<br />

precipitate, contributing to the stability of the entrapment. 38,39 This loading technology provides<br />

substantial stability with negligible drug leakage in circulation while still enabling satisfactory rates<br />

of drug release in tissues and malignant effusions. 40<br />

Studies in animal tumor models with doxorubicin encapsulated in pegylated and other longcirculating<br />

liposomes have demonstrated increased anti-tumor activity of liposomal drug as<br />

compared to optimal doses of free drug in various rodent models of syngeneic and human<br />

tumors and increased accumulation of liposomal drug in various transplantable mouse and<br />

human tumors as compared to free drug with a delayed peak tumor concentration and slow<br />

tissue clearance after injection of liposomal drug. 40<br />

The most valuable pharmacokinetic advantage of the Stealth liposomal delivery system is the<br />

enhancement of tumor exposure to doxorubicin as a result of the accumulation of liposomes in<br />

tumors <strong>by</strong> the EPR effect (Figure 29.2) as demonstrated in animal models and in some forms of<br />

human cancer. When the tissue uptake of PLD was examined in a couple of syngeneic mouse tumor<br />

models, it was found that tumor drug uptake linearly correlated with dose, and liver drug uptake<br />

showed a saturation profile. In the case of free doxorubicin, liver uptake linearly increased with<br />

dose, and tumor uptake marginally increased with dose. As a result, the delta of tumor drug<br />

concentration in favor of PLD was substantially greater at high doses. 15 These results suggest<br />

FIGURE 29.1 Schematic diagram of pegylated (Stealth) liposome encapsulating doxorubicin in the water<br />

phase (DOXIL, PLD). The PEG coating is critical for stabilizing and protecting the liposome surface from<br />

opsonization <strong>by</strong> plasma proteins, conferring prolonged circulation time. The gradient-driven loading of doxorubicin<br />

results in a high drug concentration and gelification in the liposome water phase, ensuring stable drug<br />

retention during circulation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


602<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 29.2 (See color insert following page 522.) Deposition of PLD in M109 carcinoma subcutaneously<br />

implanted in BALB/c mice <strong>by</strong> the EPR effect. The mouse was sacrificed 48 h after intravenous injection of<br />

20 mg/kg PLD. Left panel: note the red-orange to bluish color in the tumor and peritumoral halo, suggesting<br />

accumulation of doxorubicin. Right panel: note the strong tumor fluorescence observed with UV light,<br />

confirming the presence of large amounts of doxorubicin. The drug levels measured in doxorubicinequivalents<br />

were: tumor, 86 mg/g (i.e., w35% of injected dose per gram); peritumoral halo, 40 mg/g;<br />

normal skin, 0.5 mg/g.<br />

a passive process of liposomal uptake into tumor with non-saturable kinetics, whereas the liver<br />

uptake is mediated <strong>by</strong> receptor-mediated endocytosis and shows a saturable profile.<br />

In preclinical therapeutic studies using a variety of rodent tumors and human xenografts in<br />

immunodeficient nude mice, PLD was more effective than free doxorubicin and other (non-pegylated)<br />

formulations of liposomal doxorubicin. 11 In a few instances, the activity of PLD preparations<br />

was matched, but not surpassed, <strong>by</strong> non-pegylated preparations of liposomal doxorubicin. 1 In many<br />

of these studies, the improved efficacy of PLD was obtained at milligram-equivalent doses to the<br />

MTD of free doxorubicin, indicating that there was a net therapeutic gain per milligram drug,<br />

independently of toxicity buffering. A study addressing this issue examined the activity of escalating<br />

doses of PLD and doxorubicin against implants of the mouse 3LL tumor and concluded that<br />

the activity of 1–2 mg/kg DOXIL was approximately equivalent to 9 mg/kg doxorubicin, i.e., a<br />

6-fold enhancement in efficacy. 41 Similar observations were made in the M109 model pointing to a<br />

4-fold advantage for PLD as compared to free doxorubicin, i.e., a dose of 2.5 mg/kg PLD was at<br />

least as effective as 10 mg/kg free doxorubicin. 15<br />

There is a large body of preclinical data on other liposome formulations of anti-cancer agents<br />

moving into clinical development or already approved for clinical use. A tentative list of these<br />

formulations is presented in Table 29.2. Obviously, in some cases, the added value of these<br />

formulations has not been or will not be sufficient to justify further development despite positive<br />

preclinical data.<br />

29.7 LIPOSOMAL ANTHRACYCLINES IN THE CLINIC—DOXIL (PLD)<br />

The anthracycline antibiotic doxorubicin has a broad spectrum of antineoplastic action and a<br />

correspondingly widespread degree of clinical use. In addition to its role in the treatment of<br />

breast cancer, doxorubicin is indicated in the treatment of various cancers of the lymphatic and<br />

hematopoyetic systems, gastric carcinoma, small-cell cancer of the lung, soft tissue, and bone<br />

sarcomas as well as cancer of the uterus, ovary, bladder, and thyroid. Unfortunately, toxicity<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 603<br />

TABLE 29.2<br />

Formulations of Liposomal Anti-Cancer Agents Clinically Tested a<br />

Pegylated Non-Pegylated Regional Therapy Non-Cytotoxic<br />

DOXIL (PLD) a<br />

PLD with DSPC 74,b<br />

MCC-465<br />

Immunoliposome 78<br />

SPI-077 63<br />

Lipoplatin 84<br />

Myocet 71<br />

DaunoXome 75<br />

Liposomal Annamycin 79<br />

DepoCyt (intrathecal) 72<br />

L-NDDP (intrapleural) 76<br />

Liposomal Camptothecin<br />

(aerosol) 80<br />

L-MTP-PE 73<br />

Liposomal ATRA 77<br />

BLP25 Liposomal Vaccine 81<br />

Liposomal Vincristine 57<br />

Liposomal IL2 (aerosol) 82<br />

Liposomal antisense ODN 83<br />

Lurtotecan/OSI-211 56,85<br />

Liposomal E1A (intratumoral)<br />

86<br />

S-CKD602 87,c<br />

OSI-7904L (TS inhibitor) 88<br />

LE-Paclitaxel 58<br />

a<br />

This list refers to formulations tested in the last 10 years and does not intend to be comprehensive given the difficulties<br />

involved in tracking all the relevant publications.<br />

b<br />

This formulation is from Taiwan and differs from DOXIL in that DSPC is the main lipid component.<br />

c<br />

Published information on this formulation of a topo I inhibitor is available only in abstract form. The reference given here<br />

refers to a preclinical study done with a similar formulation.<br />

often limits the therapeutic activity of doxorubicin and may preclude adequate dosing. Other<br />

common complications of conventional anthracycline therapy include alopecia and dose-limiting<br />

myelosuppression. Most importantly, cardiotoxicity limits the cumulative dose of conventional<br />

anthracycline that can be given safely. 42<br />

Encapsulation of anthracyclines within liposomes significantly alters the drug pharmacokinetic<br />

profile with a tendency to selective enhancement of drug concentration in tumors. 43 In animal<br />

studies, these pharmacologic effects resulted in maintained or enhanced anthracycline efficacy and<br />

safety in a variety of experimental tumor types. 44 Improved safety and/or therapeutic profile of<br />

liposomal anthracycline therapy in KS, ovarian cancer, breast cancer, and multiple myeloma have<br />

been reported. 45 The relative lack of cardiotoxicity with liposomal anthracycline therapy is an<br />

important asset of the liposomal approach. 42 Therefore, in most instances where conventional<br />

anthracycline therapy is effective but the required course of treatment could lead to a high risk<br />

of toxicity, the use of liposomal anthracyclines should be preferred.<br />

The most commonly used liposomal anthracycline formulation in the clinic is DOXIL<br />

(DOXIL w , Caelyx w , PLD). PLD embodies the pharmacokinetic advantages of stealth liposomes<br />

(Figure 29.1). The PLD formulation consists of doxorubicin encapsulated in pegylated liposomes<br />

formulated with a hydrogenated soybean PC and cholesterol. 40 PLD was granted initial market<br />

clearance in 1995 <strong>by</strong> the U.S. Food and Drug Administration (FDA) for use in treatment of AIDSrelated<br />

KS. In 1999, PLD was granted U.S. market clearance for use in the treatment of recurrent<br />

metastatic carcinoma of the ovary based on a superior therapeutic index over topotecan, the<br />

comparator drug. In January 2003, the European Commission of the European Union granted<br />

centralized marketing authorization to PLD as monotherapy for metastatic breast cancer based<br />

on equivalent efficacy to doxorubicin and reduced cardiac toxicity. In addition, a phase III trial<br />

recently completed in the U.S. and Europe investigating the safety and efficacy of PLD as a<br />

replacement for doxorubicin in multiple myeloma found equal efficacy and reduced toxicity for<br />

PLD treatment, paving the way for PLD to become standard therapy in multiple myeloma. 46<br />

PLD was recognized 10 years ago as a liposomal doxorubicin formulation with unique pharmacokinetics<br />

and had a major change in the clinical toxicity profile. Clinical pharmacokinetic<br />

studies have indicated that PLD dramatically prolongs the circulation time of doxorubicin<br />

in agreement with preclinical studies. In 1994, the results of a pharmacokinetic study where<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


604<br />

Nanotechnology for Cancer Therapy<br />

15 patients were sequentially given the same dose in drug-equivalents of free doxorubicin and PLD<br />

were published. 47 A dramatic reduction in the drug clearance and volume of distribution resulting in<br />

a 1000-fold increase in the area under the curve (AUC) was observed with the liposomal formulation.<br />

It was also found that practically all the drug circulating in plasma is in liposomeencapsulated<br />

form. Metabolites in plasma were undetectable or were present at very low levels.<br />

However, they were readily detected in urine 24 hours or later after injection, indicating that the<br />

drug had become bioavailable. The following drug distribution picture had emerged from this<br />

initial study and more recent ones: 40<br />

1. Drug circulates in plasma for prolonged periods of time (half-life in the range of<br />

w50–80 hours) in liposome-encapsulated form. Despite its prolonged presence in<br />

blood, the drug is not bioavailable as long as it remains in the interior of a<br />

circulating liposome.<br />

2. Most of the injected drug (O95%) is distributed to tissues in liposome-encapsulated<br />

form. Once in tissues, drug leakage and liposome breakdown with or without liposome<br />

internalization <strong>by</strong> cells gradually provides a pool of bioavailable drug.<br />

3. The rate of metabolite production is slower than the rate of renal clearance of metabolites.<br />

As a result, metabolites do not accumulate in plasma but can be detected in urine.<br />

4. A small fraction of injected drug (!5%) leaks from circulating liposomes and is handled<br />

as free drug with fast plasma clearance and rapid metabolism. This drug fraction is the<br />

source of small amounts of metabolites that can be sometimes detected in plasma.<br />

In 1995, a phase I study of PLD in patients with solid tumors 48 provided clear evidence of a<br />

major change in the toxicity profile with muco-cutaneous toxicities as the major dose-limiting<br />

toxicities. In contrast, myelosuppression and alopecia were minor and cardiotoxicity was conspicuously<br />

absent. The maximal tolerated dose was established as 60 mg/m 2 with mucositis being the<br />

dose-limiting toxicity. It was also found that the optimal dosing interval for retreatment was four<br />

weeks, rather than the standard 3-week schedule of doxorubicin. The dose-schedule limiting<br />

toxicity was a form of skin toxicity known as hand-foot syndrome, also referred to as palmarplantar<br />

erythrodysesthesia (PPE), that appears to be related to the long half-life of PLD. Therefore,<br />

it became well-established that the PLD formulation imparts a significant pharmacokinetic–pharmacodynamic<br />

change to the drug doxorubicin, unprecedented in magnitude for any intravenous<br />

drug delivery system. Later, data gathered from further clinical studies in metastatic breast cancer<br />

and recurrent ovarian cancer brought down the recommended dose of PLD from 60 to 40–50 mg/m 2<br />

once every four weeks. 49 This dose reduction was needed mainly to prevent skin toxicity, resulting<br />

from successive courses of therapy.<br />

KS is a multi-focal tumor affecting skin and sometimes mucosas well known for its very high<br />

microvascular permeability. Profuse extravasation of colloidal gold-labeled Stealth liposomes in a<br />

transgenic mouse model of KS has been shown. 50 In AIDS patients, KS is frequent and has an<br />

aggressive course. Therefore, this condition was chosen for initial clinical testing of PLD in phase<br />

II-III studies. Indeed, PLD as single agent therapy demonstrated a significantly greater efficacy and<br />

better safety than standard chemotherapy (combinations of bleomycin and vincristine with or<br />

without doxorubicin) and was also effective as second line chemotherapy in pretreated patients.<br />

Because of the extreme sensitivity of KS to chemotherapy, a low and relatively subtoxic dose of<br />

20 mg/m 2 every three weeks is sufficient for effective treatment. 51<br />

Further to the successful application in various malignant conditions, there was another major<br />

development in the clinical research with PLD related to the prevention of the cardiac toxicity of<br />

doxorubicin. Evaluation of the cardiac function in patients receiving PLD revealed a major risk<br />

reduction of cardiotoxicity as compared to free doxorubicin historical data. A retrospective analysis<br />

of patients treated with large cumulative doses of PLD did not reveal any significant cardiac toxicity<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 605<br />

despite the fact that some of these patients were treated with three times as much as the maximal<br />

cumulative dose acceptable for free doxorubicin. 52 Two additional reports focusing on cardiac<br />

biopsies of PLD-treated patients at high cumulative doses demonstrated minimal pathology findings<br />

in line with the lack of clinical cardiotoxicity. 53,54 The cardiac safety of PLD was confirmed in<br />

a phase III study in metastatic breast cancer, showing a drastic reduction in cardiotoxicity in PLDtreated<br />

patients in comparison to free doxorubicin treatment. 55<br />

29.8 DEVELOPMENT OF OTHER LIPOSOME-ENTRAPPED<br />

CYTOTOXIC AGENTS<br />

Besides anthracyclines, other cytotoxic compounds currently being developed in liposomes<br />

include: Topoisomerase I inhibitory properties such as camptothecin analogs, 56 mitotic spindle<br />

poisons such as Vinca alkaloids and taxanes, 57,58 and DNA-reactive agents such as cisplatin, 59 and<br />

mitomycin C (MMC). 60 For cell cycle phase-specific cytotoxic drugs of the first two groups (e.g.,<br />

topotecan, vinorelbine), the anti-tumor activity tends to increase with liposome encapsulation as the<br />

time of exposure is extended <strong>by</strong> exploiting the liposomal slow release features. An additional<br />

advantage for liposome encapsulation of camptothecin analogs derives from their improved stability<br />

in the mildly acidic environment of the liposome water phase as opposed to the instability of<br />

their active lactone configuration when present in free form at pH 7.4 in plasma. For cisplatin, a<br />

major advantage of liposome encapsulation would be reduction of the nephrotoxicity. At this time,<br />

none of these compounds have yet been approved <strong>by</strong> a regulatory agency and none has gone<br />

through phase III studies.<br />

It should be noted that the development of liposomal formulations of cytotoxic agents has often<br />

failed for different reasons. The formulation of a liposomal anti-cancer agent is a complex process<br />

with at least three distinct variables that may affect the outcome and the risk of failure: the design of<br />

the liposome carrier, the choice of the drug, and the method of drug encapsulation. One example<br />

illustrating the complexity of the relationship between formulation and pharmacological effects is<br />

provided <strong>by</strong> a product known as SPI-77, consisting of cisplatin encapsulated in pegylated liposomes.<br />

These liposomes have long-circulating characteristics, retain the drug in plasma<br />

exceedingly well, reduce drug toxicity, and produce a high and long-lasting accumulation of<br />

drug in implanted tumors. 61 However, because the Pt exposure was measured in tumor extracts,<br />

it is unclear whether the Pt measured was SPI-77 (i.e., liposomally encapsulated Pt), protein-bound<br />

Pt, or unbound Pt. The anti-tumor activity of SPI-77 in preclinical models was variable and, in some<br />

cases, inferior <strong>by</strong> comparison to free drug. 61 In clinical studies, SPI-77 was inactive and showed no<br />

dose-limiting toxicities even at doses greater than 2-fold the MTD of free cisplatin. 62 It turned out<br />

that cisplatin release from these liposomes was minimal in in vitro 61 as well as in vivo systems as<br />

indicated <strong>by</strong> the low occurrence of DNA adducts, resulting in a major reduction of bioavailability. 63<br />

Indeed, microdialysis experiments in animal tumor models suggest that the amount of unbound Pt<br />

released <strong>by</strong> SPI-77 in tumor extracellular fluid is significantly lower when compared with free<br />

cisplatin. 64 The lack of activity of SPI-77 in early-phase clinical studies has led to the discontinuation<br />

of its further clinical development.<br />

In some instances, a strategy based on chemically modifying the drug is needed to achieve<br />

stable encapsulation. It is a formidable challenge to develop liposome formulations that retain the<br />

drug payload in the blood stream during prolonged periods of circulation, there<strong>by</strong> taking full<br />

advantage of the pharmacokinetic and biodistribution benefits of the pegylated liposomal carriers. 40<br />

Passive encapsulation of MMC in liposomes and even encapsulation of cyclodextrin–MMC<br />

complexes in liposomes did not result in stable formulations as a result of the rapid leakage of<br />

MMC. Recently, an alternative approach to delivering MMC via long-circulating liposomes has<br />

been proposed based on the design of a lipophilic prodrug with excellent retention in liposomes, but<br />

gradually releasing active MMC upon exposure to appropriate environmental conditions. 60<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


606<br />

Nanotechnology for Cancer Therapy<br />

Attachment of drugs to bilayer-compatible lipids can ensure prolonged association with a liposome.<br />

65,66 The MMC prodrug consists of a conjugate of MMC attached to 1,2-distearoyl glycerol<br />

lipid via a cleavable dithiobenzyl linker. Upon thiolytic cleavage of the disulfide-substituted benzyl<br />

urethane, MMC is released and becomes bioavailable. Intravenous administration of a pegylated<br />

liposomal formulation of the MMC prodrug in rats resulted in long circulation time and slow<br />

clearance consistent with Stealth pharmacokinetics, indicating stable prodrug retention in liposomes.<br />

The therapeutic index of this liposomal formulation was superior to that of free MMC in<br />

three animal tumor models, including a mouse tumor model with a multidrug-resistant phenotype<br />

where PLD was essentially inactive. The potential added value of this approach is 2-fold: effective<br />

treatment of multidrug- resistant tumors while significantly buffering the toxicity of MMC.<br />

29.9 FUTURE AVENUES OF LIPOSOME DEVELOPMENT—RECEPTOR<br />

TARGETED LIPOSOMES<br />

As liposomes remain one of the most attractive platforms for systemic drug delivery, an increased<br />

expansion and sophistication of these systems would be expected in the forthcoming future. The<br />

most logical improvement is the coupling of a ligand to the surface of the liposome that will target<br />

the vesicles to a specific cell-surface receptor followed <strong>by</strong> internalization and intracellular delivery<br />

of the liposome drug cargo. Examples in this direction are the targeting of PLD to HER2-expressing<br />

or folate-receptor expressing cancer cells using, respectively, a specific anti-HER2 scFv 67 or a<br />

folate conjugate anchored to the liposome surface. 30 Another example is the tumor vascular<br />

targeting of liposomes with endothelium-specific peptides associated to liposomes. 68 Amajor<br />

advantage of targeted liposomal nanocarriers over ligand-drug bioconjugates is the delivery-amplifying<br />

effect of the former that sometimes can provide to the target cell a ratio of 1,000 drug<br />

molecules per single ligand–receptor interaction.<br />

One important technological and conceptual advance in this field is the ability to insert ligands<br />

onto preformed liposomes using lipophilic tails as anchors into the bilayer. Ligand post-insertion<br />

(as opposed to pre-insertion during liposome formation) has been demonstrated for antibody fragments<br />

69 and small ligands such as folate. 30 This approach enables, on the one hand, the possibility<br />

of using end-product formulations without interfering with the pharmaceutical methods of preparation,<br />

and, on the other hand, the flexibility of ligand choice, depending on the target to be<br />

addressed. One encouraging observation, at least for folate-targeted liposomes, is that pegylated<br />

liposomes are able to retain during in vivo circulation and extravasation to malignant ascites the<br />

targeting ligand and the ability to bind to target cells, suggesting that these systems are stable<br />

enough for in vivo applications, requiring prolonged circulation before encountering the target. 70<br />

Except for rare instances, tumor cells are not directly exposed to the blood stream. Therefore,<br />

for an intra-vascular targeting device to access the targeted tumor cell receptor, it has first to cross<br />

the vascular endothelium and diffuse into the interstitial fluid. Experimental data with antibodytargeted<br />

liposomes and folate-targeted liposomes indicate that liposome deposition in tumors is<br />

similar for both targeted and non-targeted systems, 29,70 supporting the hypothesis that extravasation<br />

is the rate-limiting step of liposome accumulation in tumors. However, once liposomes have<br />

penetrated the tumor interstitial fluid, binding of targeted liposomes to the targeted receptor may<br />

occur, shifting the intra-tumor distribution from the extracellular compartment to the tumor cell<br />

compartment as recently shown for a mouse ascitic tumor. 70 Retrograde movement of liposomes<br />

into the blood stream, if any, will be reduced for liposomes with binding affinity to a tumor cell<br />

receptor. Therefore, the theoretical advantages of a targeted liposomal drug delivery system over a<br />

non-targeted one are primarily derived from a relative shift of liposome distribution from the<br />

extracellular fluid to the tumor cell compartment, and, possibly, from prolonged liposome retention<br />

in tumor. On the negative side, the main disadvantage of a targeted system to a cancer cell receptor<br />

is the difficulty of a large nanosize assembly such as liposomes to penetrate a solid tumor mass,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 607<br />

confronted <strong>by</strong> the high interstitial fluid pressure that often characterizes tumor masses at clinically<br />

detectable size. 31 One such system targeted to the Her2 receptor has shown superior therapeutic<br />

efficacy over its non-targeted counterpart. With the maturation of the pharmaceutical technology<br />

required for the application of targeted liposomal drug delivery systems, the final word regarding<br />

their added value over non-targeted systems awaits clinical trials.<br />

29.10 CONCLUSIONS<br />

Despite huge research efforts, with liposomal drugs is cancer therapy anthracycline formulations,<br />

and specifically, the leading compound, PLD, are the only liposome products that have been able,<br />

so far, to fill a niche in the anti-cancer armamentarium. As this chapter has discussed, this is due, at<br />

least in part, to the complexity of the pharmaceutical technology involved that requires drugspecific<br />

solutions and has a direct impact on the pharmacology of the end compound. Many of<br />

the oncology opinion leaders regard liposome formulations, with liposomal in cancer therapy drugs<br />

as a cosmetic improvement of toxic drugs that will become obsolete in the near future once smart<br />

new drugs are developed, given the increasing knowledge of the molecular biology of cancer cells.<br />

However, a realistic scenario favored <strong>by</strong> the majority of clinicians predicts that the use of broadspectrum<br />

cytotoxic drugs will remain as a major tool in cancer therapy for a few more decades to<br />

come. Moreover, the nanomedicine approach is perceived today as a key to future breakthroughs<br />

because of the complexity of living systems. These factors may reinvigorate the interest in particulate<br />

drug delivery systems such as liposomes and augur that the use of liposome carriers in cancer<br />

therapy will expand in the foreseeable future.<br />

REFERENCES<br />

1. Gabizon, A. A., Pegylated liposomal doxorubicin: Metamorphosis of an old drug into a new form of<br />

chemotherapy, Cancer Invest., 19(4), 424–436, 2001.<br />

2. Gradishar, W. J., Tjulandin, S., Davidson, N., Shaw, H., Desai, N., Bhar, P. et al., Phase III trial of<br />

nanoparticle albumin-bound paclitaxel compared with polyethylated castor oil-based paclitaxel in<br />

women with breast cancer, J. Clin. Oncol., 23(31), 7794–7803, 2005.<br />

3. Lasic, D. D. and Papahadjopoulos, D., Medical Applications of Liposomes, Elsevier, Amsterdam,<br />

1998.<br />

4. Barenholz, Y., Relevancy of drug loading to liposomal formulation therapeutic efficacy, J. Liposome<br />

Res., 13(1), 1–8, 2003.<br />

5. Moghimi, S. M., Hunter, A. C., and Murray, J. C., Nanomedicine: Current status and future prospects,<br />

FASEB J., 19(3), 311–330, 2005.<br />

6. Hanahan, D. and Weinberg, R. A., The hallmarks of cancer, Cell, 100(1), 57–70, 2000.<br />

7. Van Den Bossche, B. and Van de Wiele, C., Receptor imaging in oncology <strong>by</strong> means of nuclear<br />

medicine: Current status, J. Clin. Oncol., 22(17), 3593–3607, 2004.<br />

8. Kastan, M. B., Molecular biology of cancer: The cell cycle, In Cancer: Principles & Practice of<br />

Oncology, DeVita, V. T., Hellman, S., and Rosenberg, S. A., Eds., 5th ed., Lippincott-Raven, Philadelphia,<br />

pp. 121–134, 1997.<br />

9. Jain, R. K., Delivery of molecular medicine to solid tumors: Lessons from in vivo imaging of gene<br />

expression and function, J. Control. Release, 74(1-3), 7–25, 2001.<br />

10. Papahadjopoulos, D., Allen, T. M., Gabizon, A., Mayhew, E., Matthay, K., Huang, S. K. et al.,<br />

Sterically stabilized liposomes: Improvements in pharmacokinetics and antitumor therapeutic efficacy,<br />

Proc. Natl Acad. Sci. USA, 88(24), 11460–11464, 1991.<br />

11. Gabizon, A. and Martin, F., Polyethylene glycol-coated (pegylated) liposomal doxorubicin. Rationale<br />

for use in solid tumours, Drugs, 54(Suppl. 4), 15–21, 1997.<br />

12. Swabb, E. A., Wei, J., and Gullino, P. M., Diffusion and convection in normal and neoplastic tissues,<br />

Cancer Res., 34(10), 2814–2822, 1974.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


608<br />

Nanotechnology for Cancer Therapy<br />

13. Harrington, K. J., Rowlinson-Busza, G., Syrigos, K. N., Abra, R. M., Uster, P. S., Peters, A. M. et al.,<br />

Influence of tumour size on uptake of 111 ln-DTPA-labelled pegylated liposomes in a human tumour<br />

xenograft model, Br. J. Cancer, 83(5), 684–688, 2000.<br />

14. Safra, T., Groshen, S., Jeffers, S., Tsao-Wei, D. D., Zhou, L., Muderspach, L. et al., Treatment of<br />

patients with ovarian carcinoma with pegylated liposomal doxorubicin: Analysis of toxicities and<br />

predictors of outcome, Cancer, 91(1), 90–100, 2001.<br />

15. Gabizon, A., Tzemach, D., Mak, L., Bronstein, M., and Horowitz, A. T., Dose dependency of<br />

pharmacokinetics and therapeutic efficacy of pegylated liposomal doxorubicin (DOXIL) in murine<br />

models, J. Drug Target, 10(7), 539–548, 2002.<br />

16. Gregoriadis, G., Leathwood, P. D., and Ryman, B. E., Enzyme entrapment in liposomes, FEBS Lett.,<br />

14(2), 95–99, 1971.<br />

17. Allen, T. M. and Chonn, A., Large unilamellar liposomes with low uptake into the reticuloendothelial<br />

system, FEBS Lett., 223(1), 42–46, 1987.<br />

18. Gabizon, A. and Papahadjopoulos, D., Liposome formulations with prolonged circulation time in<br />

blood and enhanced uptake <strong>by</strong> tumors, Proc. Natl Acad. Sci. USA, 85(18), 6949–6953, 1988.<br />

19. Allen, T. M., Austin, G. A., Chonn, A., Lin, L., and Lee, K. C., Uptake of liposomes <strong>by</strong> cultured<br />

mouse bone marrow macrophages: Influence of liposome composition and size, Biochim. Biophys.<br />

Acta, 1061(1), 56–64, 1991.<br />

20. Woodle, M. C. and Lasic, D. D., Sterically stabilized liposomes, Biochim. Biophys. Acta, 1113(2),<br />

171–199, 1992.<br />

21. Klibanov, A. L., Maruyama, K., Torchilin, V. P., and Huang, L., Amphipathic polyethyleneglycols<br />

effectively prolong the circulation time of liposomes, FEBS Lett., 268(1), 235–237, 1990.<br />

22. Allen, T. M., Hansen, C., Martin, F., Redemann, C., and Yau-Young, A., Liposomes containing<br />

synthetic lipid derivatives of poly(ethylene glycol) show prolonged circulation half-lives in vivo,<br />

Biochim. Biophys. Acta, 1066(1), 29–36, 1991.<br />

23. Lasic, D. D. and Martin, F. J., Stealth Liposomes, CRC Press, Boca Raton, FL, 1995.<br />

24. Huang, S. K., Lee, K. D., Hong, K., Friend, D. S., and Papahadjopoulos, D., Microscopic localization<br />

of sterically stabilized liposomes in colon carcinoma-bearing mice, Cancer Res., 52(19), 5135–5143,<br />

1992.<br />

25. Yuan, F., Leunig, M., Huang, S. K., Berk, D. A., Papahadjopoulos, D., and Jain, R. K., Microvascular<br />

permeability and interstitial penetration of sterically stabilized (stealth) liposomes in a human tumor<br />

xenograft, Cancer Res., 54(13), 3352–3356, 1994.<br />

26. Gabizon, A. A., Selective tumor localization and improved therapeutic index of anthracyclines<br />

encapsulated in long-circulating liposomes, Cancer Res., 52(4), 891–896, 1992.<br />

27. Bally, M. B., Masin, D., Nayar, R., Cullis, P. R., and Mayer, L. D., Transfer of liposomal drug carriers<br />

from the blood to the peritoneal cavity of normal and ascitic tumor-bearing mice, Cancer Chemother.<br />

Pharmacol., 34(2), 137–146, 1994.<br />

28. Maeda, H., The enhanced permeability and retention (EPR) effect in tumor vasculature: The key role<br />

of tumor-selective macromolecular drug targeting, Adv. Enzyme Regul., 41, 189–207, 2001.<br />

29. Goren, D., Horowitz, A. T., Zalipsky, S., Woodle, M. C., Yarden, Y., and Gabizon, A., Targeting of<br />

stealth liposomes to erbB-2 (Her/2) receptor: In vitro and in vivo studies, Br. J. Cancer, 74(11),<br />

1749–1756, 1996.<br />

30. Gabizon, A., Shmeeda, H., Horowitz, A. T., and Zalipsky, S., Tumor cell targeting of liposomeentrapped<br />

drugs with phospholipid-anchored folic acid-PEG conjugates, Adv. Drug Deliv. Rev., 56(8),<br />

1177–1192, 2004.<br />

31. Stohrer, M., Boucher, Y., Stangassinger, M., and Jain, R. K., Oncotic pressure in solid tumors is<br />

elevated, Cancer Res., 60(15), 4251–4255, 2000.<br />

32. Harrington, K. J., Mohammadtaghi, S., Uster, P. S., Glass, D., Peters, A. M., Vile, R. G. et al.,<br />

Effective targeting of solid tumors in patients with locally advanced cancers <strong>by</strong> radiolabeled pegylated<br />

liposomes, Clin. Cancer Res., 7(2), 243–254, 2001.<br />

33. Symon, Z., Peyser, A., Tzemach, D., Lyass, O., Sucher, E., Shezen, E. et al., Selective delivery of<br />

doxorubicin to patients with breast carcinoma metastases <strong>by</strong> stealth liposomes, Cancer, 86(1), 72–78,<br />

1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 609<br />

34. Northfelt, D. W., Martin, F. J., Working, P., Volberding, P. A., Russell, J., Newman, M. et al.,<br />

Doxorubicin encapsulated in liposomes containing surface-bound polyethylene glycol: Pharmacokinetics,<br />

tumor localization, and safety in patients with AIDS-related Kaposi’s sarcoma, J. Clin.<br />

Pharmacol., 36(1), 55–63, 1996.<br />

35. Young, R. C., Ozols, R. F., and Myers, C. E., The anthracycline antineoplastic drugs, N. Engl. J. Med.,<br />

305(3), 139–153, 1981.<br />

36. Forssen, E. A. and Tokes, Z. A., In vitro and in vivo studies with adriamycin liposomes, Biochem.<br />

Biophys. Res. Commun., 91(4), 1295–1301, 1979.<br />

37. Gabizon, A. A., Liposomal anthracyclines, Hematol. Oncol. Clin. North Am., 8(2), 431–450, 1994.<br />

38. Lasic, D. D., Frederik, P. M., Stuart, M. C., Barenholz, Y., and McIntosh, T. J., Gelation of liposome<br />

interior. A novel method for drug encapsulation, FEBS Lett., 312(2–3), 255–258, 1992.<br />

39. Haran, G., Cohen, R., Bar, L. K., and Barenholz, Y., Transmembrane ammonium sulfate gradients in<br />

liposomes produce efficient and stable entrapment of amphipathic weak bases, Biochim. Biophys.<br />

Acta, 1151(2), 201–215, 1993.<br />

40. Gabizon, A., Shmeeda, H., and Barenholz, Y., Pharmacokinetics of pegylated liposomal Doxorubicin:<br />

Review of animal and human studies, Clin. Pharmacokinet., 42(5), 419–436, 2003.<br />

41. Colbern, G. T., Hiller, A. J., Musterer, R. S. et al., Significant increase in antitumor potency of<br />

doxorubicin HCI <strong>by</strong> its encapsulation in pegylated liposomes, J. Liposome Res., 9, 523–538, 1999.<br />

42. Ewer, M. S., Martin, F. J., Henderson, C., Shapiro, C. L., Benjamin, R. S., and Gabizon, A. A., Cardiac<br />

safety of liposomal anthracyclines, Semin. Oncol., 31(6 Suppl. 13), 161–181, 2004.<br />

43. Allen, T. M. and Martin, F. J., Advantages of liposomal delivery systems for anthracyclines, Semin.<br />

Oncol., 31(6 Suppl. 13), 5–15, 2004.<br />

44. Vail, D. M., Amantea, M. A., Colbern, G. T., Martin, F. J., Hilger, R. A., and Working, P. K.,<br />

Pegylated liposomal doxorubicin: Proof of principle using preclinical animal models and pharmacokinetic<br />

studies, Semin. Oncol., 31(6 Suppl. 13), 16–35, 2004.<br />

45. Henderson, C. I., Established and emerging roles of liposomal anthracyclines in oncology. Introduction,<br />

Semin. Oncol., 31(6 Suppl. 13), 1–4, 2004.<br />

46. Rifkin, R. M., Gregory, S. A., Mohrbacher, A., and Hussein, M. A., PEGylated liposomal doxorubicin,<br />

vincristine, and dexamethasone provide significant reduction in toxicity compared with<br />

doxorubicin, vincristine, and dexamethasone in patients with newly diagnosed multiple myeloma,<br />

Cancer, 106(4), 848–858, <strong>2006</strong>.<br />

47. Gabizon, A., Catane, R., Uziely, B., Kaufman, B., Safra, T., Cohen, R. et al., Prolonged circulation<br />

time and enhanced accumulation in malignant exudates of doxorubicin encapsulated in polyethyleneglycol<br />

coated liposomes, Cancer Res., 54(4), 987–992, 1994.<br />

48. Uziely, B., Jeffers, S., Isacson, R., Kutsch, K., Wei-Tsao, D., Yehoshua, Z. et al., Liposomal doxorubicin:<br />

Antitumor activity and unique toxicities during two complementary phase I studies, J. Clin.<br />

Oncol., 13(7), 1777–1785, 1995.<br />

49. Alberts, D. S., Muggia, F. M., Carmichael, J., Winer, E. P., Jahanzeb, M., Venook, A. P. et al.,<br />

Efficacy and safety of liposomal anthracyclines in phase I/II clinical trials, Semin. Oncol., 31(6<br />

Suppl. 13), 53–90, 2004.<br />

50. Huang, S. K., Martin, F. J., Jay, G., Vogel, J., Papahadjopoulos, D., and Friend, D. S., Extravasation<br />

and transcytosis of liposomes in Kaposi’s sarcoma-like dermal lesions of transgenic mice bearing the<br />

HIV tat gene, Am. J. Pathol., 143(1), 10–14, 1993.<br />

51. Krown, S. E., Northfelt, D. W., Osoba, D., and Stewart, J. S., Use of liposomal anthracyclines in<br />

Kaposi’s sarcoma, Semin. Oncol., 31(6 Suppl. 13), 36–52, 2004.<br />

52. Safra, T., Muggia, F., Jeffers, S., Tsao-Wei, D. D., Groshen, S., Lyass, O. et al., Pegylated liposomal<br />

doxorubicin (doxil): Reduced clinical cardiotoxicity in patients reaching or exceeding cumulative<br />

doses of 500 mg/m 2 , Ann. Oncol., 11(8), 1029–1033, 2000.<br />

53. Berry, G., Billingham, M., Alderman, E., Richardson, P., Torti, F., Lum, B. et al., The use of cardiac<br />

biopsy to demonstrate reduced cardiotoxicity in AIDS Kaposi’s sarcoma patients treated with pegylated<br />

liposomal doxorubicin, Ann. Oncol., 9(7), 711–716, 1998.<br />

54. Gabizon, A. A., Lyass, O., Berry, G. J., and Wildgust, M., Cardiac safety of pegylated liposomal<br />

doxorubicin (Doxil/Caelyx) demonstrated <strong>by</strong> endomyocardial biopsy in patients with advanced<br />

malignancies, Cancer Invest., 22(5), 663–669, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


610<br />

Nanotechnology for Cancer Therapy<br />

55. O’Brien, M. E., Wigler, N., Inbar, M., Rosso, R., Grischke, E., Santoro, A. et al., Reduced cardiotoxicity<br />

and comparable efficacy in a phase III trial of pegylated liposomal doxorubicin HCl<br />

(CAELYX/Doxil) versus conventional doxorubicin for first-line treatment of metastatic breast<br />

cancer, Ann. Oncol., 15(3), 440–449, 2004.<br />

56. Kehrer, D. F., Bos, A. M., Verweij, J., Groen, H. J., Loos, W. J., Sparreboom, A. et al., Phase I and<br />

pharmacologic study of liposomal lurtotecan, NX 211: Urinary excretion predicts hematologic<br />

toxicity, J. Clin. Oncol., 20(5), 1222–1231, 2002.<br />

57. Gelmon, K. A., Tolcher, A., Diab, A. R., Bally, M. B., Embree, L., Hudon, N. et al., Phase I study of<br />

liposomal vincristine, J. Clin. Oncol., 17(2), 697–705, 1999.<br />

58. Soepenberg, O., Sparreboom, A., de Jonge, M. J., Planting, A. S., de Heus, G., Loos, W. J. et al., Realtime<br />

pharmacokinetics guiding clinical decisions; phase I study of a weekly schedule of liposome<br />

encapsulated paclitaxel in patients with solid tumours, Eur. J. Cancer, 40(5), 681–688, 2004.<br />

59. Boulikas, T., Stathopoulos, G. P., Volakakis, N., and Vougiouka, M., Systemic lipoplatin infusion<br />

results in preferential tumor uptake in human studies, Anticancer Res., 25(4), 3031–3039, 2005.<br />

60. Gabizon, A., Tzemach, D., Horowitz, A. T., Shmeeda, H., Yeh, Y., and Zalipsky, S., Reduced toxicity<br />

and superior therapeutic activity of a Mitomycin C lipid-based prodrug incorporated in Pegylated<br />

liposomes. Clin. Cancer Res. 12(6), 1913–1920, <strong>2006</strong>.<br />

61. Bandak, S., Goren, D., Horowitz, A., Tzemach, D., and Gabizon, A., Pharmacological studies of<br />

cisplatin encapsulated in long-circulating liposomes in mouse tumor models, Anticancer Drugs,<br />

10(10), 911–920, 1999.<br />

62. Harrington, K. J., Lewanski, C. R., Northcote, A. D., Whittaker, J., Wellbank, H., Vile, R. G. et al.,<br />

Phase I-II study of pegylated liposomal cisplatin (SPI-077) in patients with inoperable head and neck<br />

cancer, Ann. Oncol., 12(4), 493–496, 2001.<br />

63. Meerum Terwogt, J. M., Groenewegen, G., Pluim, D., Maliepaard, M., Tibben, M. M., Huisman, A.<br />

et al., Phase I and pharmacokinetic study of SPI-77, a liposomal encapsulated dosage form of<br />

cisplatin, Cancer Chemother. Pharmacol., 49(3), 201–210, 2002.<br />

64. Zamboni, W. C., Gervais, A. C., Egorin, M. J., Schellens, J. H., Zuhowski, E. G., Pluim, D. et al.,<br />

Systemic and tumor disposition of platinum after administration of cisplatin or STEALTH liposomalcisplatin<br />

formulations (SPI-077 and SPI-077 B103) in a preclinical tumor model of melanoma,<br />

Cancer Chemother. Pharmacol., 53(4), 329–336, 2004.<br />

65. Mori, A., Kennel, S. J., van Borssum Waalkes, M., Scherphof, G. L., and Huang, L., Characterization<br />

of organ-specific immunoliposomes for delivery of 3 0 ,5 0 -O-dipalmitoyl-5-fluoro-2 0 -deoxyuridine in a<br />

mouse lung metastasis model, Cancer Chemother. Pharmacol., 35, 447–456, 1995.<br />

66. Asai, T., Kurohane, K., Shuto, S., Awano, H., Matsuda, A., Tsukada, H. et al., Antitumor activity of<br />

5 0 -O-dipalmitoylphosphatidyl 2 0 -C-cyano-2 0 -deoxy-1-b-D-arabino-pentofuranosylcytosine is<br />

enhanced <strong>by</strong> long-circulating liposomalization, Biol. Pharm. Bull., 21(7), 766–771, 1998.<br />

67. Park, J. W., Hong, K., Kirpotin, D. B., Colbern, G., Shala<strong>by</strong>, R., Baselga, J. et al., Anti-HER2<br />

immunoliposomes: Enhanced efficacy attributable to targeted delivery, Clin. Cancer Res., 8(4),<br />

1172–1181, 2002.<br />

68. Pastorino, F., Brignole, C., Marimpietri, D., Cilli, M., Gambini, C., Ribatti, D. et al., Vascular damage<br />

and anti-angiogenic effects of tumor vessel-targeted liposomal chemotherapy, Cancer Res., 63(21),<br />

7400–7409, 2003.<br />

69. Allen, T. M., Sapra, P., and Moase, E., Use of the post-insertion method for the formation of ligandcoupled<br />

liposomes, Cell Mol. Biol. Lett., 7(3), 889–894, 2002.<br />

70. Gabizon, A., Horowitz, A. T., Goren, D., Tzemach, D., Shmeeda, H., and Zalipsky, S., In vivo fate of<br />

folate-targeted polyethylene-glycol liposomes in tumor-bearing mice, Clin. Cancer Res., 9(17),<br />

6551–6559, 2003.<br />

71. Harris, L., Batist, G., Belt, R., Rovira, D., Navari, R., Azarnia, N. et al., Liposome-encapsulated<br />

doxorubicin compared with conventional doxorubicin in a randomized multicenter trial as first-line<br />

therapy of metastatic breast carcinoma, Cancer, 94(1), 25–36, 2002.<br />

72. Bomgaars, L., Geyer, J. R., Franklin, J., Dahl, G., Park, J., Winick, N. J. et al., Phase I trial of<br />

intrathecal liposomal cytarabine in children with neoplastic meningitis, J. Clin. Oncol., 22(19),<br />

3916–3921, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Applications of Liposomal Drug Delivery Systems to Cancer Therapy 611<br />

73. Meyers, P. A., Schwartz, C. L., Krailo, M., Kleinerman, E. S., Betcher, D., Bernstein, M. L. et al.,<br />

Osteosarcoma: A randomized, prospective trial of the addition of ifosfamide and/or muramyl tripeptide<br />

to cisplatin, doxorubicin, and high-dose methotrexate, J. Clin. Oncol., 23(9), 2004–2011, 2005.<br />

74. Hong, R. L. and Tseng, Y. L., Phase I and pharmacokinetic study of a stable, polyethylene-glycolated<br />

liposomal doxorubicin in patients with solid tumors: The relation between pharmacokinetic property<br />

and toxicity, Cancer, 91(9), 1826–1833, 2001.<br />

75. O’Byrne, K. J., Thomas, A. L., Sharma, R. A., DeCatris, M., Shields, F., Beare, S. et al., A phase I<br />

dose-escalating study of DaunoXome, liposomal daunorubicin, in metastatic breast cancer, Br.<br />

J. Cancer, 87(1), 15–20, 2002.<br />

76. Lu, C., Perez-Soler, R., Piperdi, B., Walsh, G. L., Swisher, S. G., Smythe, W. R. et al., Phase II study<br />

of a liposome-entrapped cisplatin analog (L-NDDP) administered intrapleurally and pathologic<br />

response rates in patients with malignant pleural mesothelioma, J. Clin. Oncol., 23(15),<br />

3495–3501, 2005.<br />

77. Ozpolat, B., Lopez-Berestein, G., Adamson, P., Fu, C. J., and Williams, A. H., Pharmacokinetics of<br />

intravenously administered liposomal all-trans-retinoic acid (ATRA) and orally administered ATRA<br />

in healthy volunteers, J. Pharm. Pharm. Sci., 6(2), 292–301, 2003.<br />

78. Matsumura, Y., Gotoh, M., Muro, K., Yamada, Y., Shirao, K., Shimada, Y. et al., Phase I and<br />

pharmacokinetic study of MCC-465, a doxorubicin (DXR) encapsulated in PEG immunoliposome,<br />

in patients with metastatic stomach cancer, Ann. Oncol., 15(3), 517–525, 2004.<br />

79. Booser, D. J., Esteva, F. J., Rivera, E., Valero, V., Esparza-Guerra, L., Priebe, W. et al., Phase II study<br />

of liposomal annamycin in the treatment of doxorubicin-resistant breast cancer, Cancer Chemother.<br />

Pharmacol., 50(1), 6–8, 2002.<br />

80. Verschraegen, C. F., Gilbert, B. E., Loyer, E., Huaringa, A., Walsh, G., Newman, R. A. et al., Clinical<br />

evaluation of the delivery and safety of aerosolized liposomal 9-nitro-20(s)-camptothecin in patients<br />

with advanced pulmonary malignancies, Clin. Cancer Res., 10(7), 2319–2326, 2004.<br />

81. Butts, C., Murray, N., Maksymiuk, A., Goss, G., Marshall, E., Soulieres, D. et al., Randomized phase<br />

IIB trial of BLP25 liposome vaccine in stage IIIB and IV non-small-cell lung cancer, J. Clin. Oncol.,<br />

23(27), 6674–6681, 2005.<br />

82. Skubitz, K. M. and Anderson, P. M., Inhalational interleukin-2 liposomes for pulmonary metastases:<br />

A phase I clinical trial, Anticancer Drugs, 11(7), 555–563, 2000.<br />

83. Rudin, C. M., Marshall, J. L., Huang, C. H., Kindler, H. L., Zhang, C., Kumar, D. et al., Delivery of a<br />

liposomal c-raf-1 antisense oligonucleotide <strong>by</strong> weekly bolus dosing in patients with advanced solid<br />

tumors: A phase I study, Clin. Cancer Res., 10(21), 7244–7251, 2004.<br />

84. Boulikas, T., Low toxicity and anticancer activity of a novel liposomal cisplatin (lipoplatin) in mouse<br />

xenografts, Oncol. Rep., 12(1), 3–12, 2004.<br />

85. Dark, G. G., Calvert, A. H., Grimshaw, R., Poole, C., Swenerton, K., Kaye, S. et al., Randomized trial<br />

of two intravenous schedules of the topoisomerase I inhibitor liposomal lurtotecan in women with<br />

relapsed epithelial ovarian cancer: A trial of the national cancer institute of Canada clinical trials<br />

group, J. Clin. Oncol., 23(9), 1859–1866, 2005.<br />

86. Yoo, G. H., Hung, M. C., Lopez-Berestein, G., LaFollette, S., Ensley, J. F., Carey, M. et al., Phase I<br />

trial of intratumoral liposome E1A gene therapy in patients with recurrent breast and head and neck<br />

cancer, Clin. Cancer Res., 7(5), 1237–1245, 2001.<br />

87. Colbern, G. T., Dykes, D. J., Engbers, C., Musterer, R., Hiller, A., Pegg, E. et al., Encapsulation of the<br />

topoisomerase I inhibitor GL147211C in PEGylated (STEALTH) liposomes: Pharmacokinetics and<br />

antitumor activity in HT29 colon tumor xenografts, Clin. Cancer Res., 4(12), 3077–3082, 1998.<br />

88. Beutel, G., Glen, H., Schoffski, P., Chick, J., Gill, S., Cassidy, J. et al., Phase I study of OSI-7904L, a<br />

novel liposomal thymidylate synthase inhibitor in patients with refractory solid tumors, Clin. Cancer<br />

Res., 11(15), 5487–5495, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


30<br />

CONTENTS<br />

Positively-Charged Liposomes<br />

for Targeting Tumor Vasculature<br />

Robert B. Campbell<br />

30.1 Introduction ................................................................................................................... 613<br />

30.2 The Microvascular Endothelium: Highly Accessible Target of Therapeutics ............ 614<br />

30.3 Arguments for Vascular Targeting Approach over Interstitial Drug Targeting.......... 616<br />

30.4 Influence of Cationic Charge on Vascular Targeting in Health and in Disease ......... 617<br />

30.5 Formulation Development: Alternatives to Routine Drug<br />

Design and Synthesis .................................................................................................... 617<br />

30.6 Interaction of Cationic Liposomes with Human Endothelial Cells ............................. 618<br />

30.7 Stealth (PEGylated) Liposome Technology: A Brief Communication ....................... 619<br />

30.8 PEGylated Cationic Liposomes: Are They Suited for<br />

Targeting Tumor Vessels? ............................................................................................ 620<br />

30.9 Cationic Liposomes Improve Anti-Cancer Therapy .................................................... 621<br />

30.10 Validation and Long-Term Tracking of Tumor Response to Therapy ....................... 623<br />

30.11 Concluding Remarks..................................................................................................... 623<br />

References .................................................................................................................................. 624<br />

30.1 INTRODUCTION<br />

In order to abolish the function of established vessels and/or suppress the formation of neovascular<br />

networks, various drug delivery vehicles have been used to target experimental therapeutics to the<br />

tumor vascular surface. Improving site-specific delivery, maximizing tumor response, minimizing<br />

unwanted side effects, and overcoming tumor heterogeneity are important considerations.<br />

Clinical strategies are developed in response to unmet clinical needs, but often at the expense of<br />

understanding how to effectively exploit notable differences between the tumor and host tissues.<br />

For instance, tumor vessels are generally more tortuous, over-express a variety of active genes, and<br />

are more permeable to circulating macromolecules (compared to the vasculature of normal tissues).<br />

On the other hand, relatively high interstitial fluid pressure and long interstitial transport distances<br />

still hinder effective delivery and deep penetration of drugs in solid tumors. 1 For this reason,<br />

although the development of interstitial delivery methods represents the focal point of the majority<br />

of research efforts, alternative treatment options are being developed and evaluated in parallel.<br />

The justification for parallel research efforts is due to several advantages of endotheliumspecific<br />

drug targeting over interstitial drug delivery. First, vascular networks in developing<br />

tumors make tumors vulnerable to treatment regardless of where the tumor is located in vivo.<br />

Second, the endothelium is vulnerable to the effects of circulating agents; this approach is especially<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

613


614<br />

Nanotechnology for Cancer Therapy<br />

useful for treating tumors given the fact that a steady flow of oxygen and nutrients (from the blood to the<br />

interstitial compartment) is essential to sustain active angiogenesis and rapid tumor growth. Third, the<br />

vast majority of physiological barriers pose little, if any, threat to vascular targeting strategies, and<br />

finally, endothelial cells are non-mutagenic. This last advantage deals specifically with the inability of<br />

endothelial cells to acquire drug resistance compared to cancer cells that are inherently more prone to<br />

mutation and drug resistance. Nonetheless, despite all of these advantages, preservation of normal<br />

vascular function must be achieved so that treatment is both efficacious and safe.<br />

The procedures for delivery of drugs to tumor vessels (not the interstitium) have yet to be<br />

optimized completely for all human organ systems. Whether the aim is to deliver bulky cytotoxic<br />

drugs or dynamic therapeutic genes, peptides, and proteins, minimizing unintentional injury to<br />

healthy organ tissues is critical. Until better quality treatments are made available, existing methods<br />

will have to be applied with a better understanding of their clinical limitations. Successful strategies<br />

need to satisfy a required set of criteria that include:<br />

† The drug should not be toxic to patients;<br />

† Repetitive administration of drug should be well tolerated;<br />

† The drug should be reasonably accessible to the intended cell target;<br />

† The drug should be undetectable <strong>by</strong> the host-immune system;<br />

† Drug exposure to primary target(s) should yield useful biological product(s); and<br />

† Biological products should be sufficient to elicit a desired clinical response.<br />

Although these guidelines represent the gold standard for a number of routine clinical<br />

procedures and given that some human tumors meagerly respond to standard treatment options<br />

(yielding inadequate to mediocre responses at best), these routine procedures are simply not<br />

sufficient for treating all human cancers.<br />

In general, the endothelium is considered a non-traditional target for FDA approved agents such<br />

as doxorubicin, vincristine, and paclitaxel; therefore, much of our understanding in connection with<br />

this approach is incomplete.<br />

Cytotoxic agents are typically systemically administered at the maximum tolerated dose<br />

(MTD). The goal has long been to deliver huge therapeutic doses to eradicate millions of viable<br />

cancer cells hidden deep inside tissues. The drugs next pass through a complicated tumor interstitial<br />

matrix to prevent the metastatic process from reaching the point of no return. Today, clinical<br />

success has been achieved with the use of some interstitial drug targeting practices. The question<br />

currently asked is: Can the use of popular MTD strategies be replaced with an approach that focuses<br />

on selective delivery of drugs to the tumor vascular compartment rather than to the interstitial tumor<br />

compartment? Moreover, will this clinical practice result in a more effective clinical response that<br />

is both efficacious and safe? To address these question and related concerns, the structural and<br />

functional role of the endothelium should be examined; identification of accessible targets along the<br />

endothelium is a good start.<br />

30.2 THE MICROVASCULAR ENDOTHELIUM: HIGHLY ACCESSIBLE TARGET<br />

OF THERAPEUTICS<br />

The endothelium is a structural barrier separating the intravascular compartment from the tissue<br />

space. Because of wide variability in anatomical structure, the vascular compartment is further<br />

grouped into three different subcategories. The subcategories are continuous, fenestrated, or<br />

discontinuous endothelia.<br />

Continuous endothelia are the most common type and are found most frequently in blood<br />

vessels lining chambers of the heart, walls of capillaries and arterioles in skeletal, skin, cardiac<br />

muscle, and connective tissue. They are well known for their relatively tight cellular junctions. 2,3<br />

This particular category of vessels is also critical in the regulation and rapid exchange of ions and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 615<br />

solutes. 2,3 Plasmalemmal vesicles involved in endothelial transport are abundant in myocardial<br />

endothelia but are far less frequently observed in capillaries of the brain. 4 The actual number of<br />

plasmalemmal vesicles existing along the continuous endothelium is heterogeneous, varying as a<br />

function of organ and tissue environment. These vesicular structures are also highly sensitive to<br />

charge characteristics, favoring associations with anionic over cationic molecules. 5<br />

Fenestrated vessels are normally found in vessels of organs that secrete (or excrete) biological<br />

fluids as in the gastrointestinal mucosa and in the glomerular capillaries of the kidneys. Fenestrae<br />

are usually between 50 and 80 nm in size, and they appear either as individual gap openings in the<br />

wall of functional vessels or as clusters. Similar to plasmalemmal vesicles, their frequency of<br />

occurrence along vessels depends on organ type and microenvironment.<br />

Often, two or more capillaries may join to form post-capillary venules. These newly formed<br />

networks are composed of a single lining of endothelial cells with a basement membrane with no<br />

smooth muscle cell attachment. These vessels are heavily involved in the exchange of molecules,<br />

and they are preferential sites of plasma extravasation as a result of the actions of vasoactive and<br />

humoral factors. Fenestrated vessels possess a negatively charged surface density as a result of high<br />

heparin sulfate proteoglycan content, and unlike plasmalemmal vesicles of continuous endothelia,<br />

they favor interaction with cationic over anionic molecules. 6<br />

Discontinuous endothelia are found primarily in the liver, spleen, and bone marrow organs. 2 In<br />

the liver sinusoids, the endothelia are not continuous, and they possess an average fenestrae size<br />

between 100 and 150 nm in diameter with the size of fenestrate often changing in response to local<br />

mediators. These changes include, but are not limited to, response to luminal pressures and potent<br />

vasodilators such as histamine and bradykinin. 7–10 An investigation into the size of vascular pore<br />

openings in animal tumor models revealed gap openings that were significantly larger than those<br />

observed along vessels in normal tissues, around 4 mm (4000 nm) in at least one tumor type but<br />

normally falling within the range of 0.4–0.6 mm (400–600 nm) for many tumor types. 11,12 Nonetheless,<br />

the evidence is overwhelmingly in favor of the development of tumor-targeted delivery of<br />

therapeutic carrier molecules that are small enough to enter through tumor vascular pores without<br />

passing through openings in normal healthy organ tissues.<br />

The endothelium is responsible for synthesizing a variety of molecules, regulating endothelial cell<br />

migration, proliferation, blood vessel maturation, and function. It has been shown to synthesize<br />

vascular growth factors, nitric oxide, collagen IV, laminin, glycosaminoglycans, and proteoglycans<br />

to highlight several proven functions. 10,13–19 The physical barrier organizes very rapidly to form<br />

monolayers, reassembles to form vascular tubes, 20 and can change specific inter- and intra-cellular<br />

signaling patterns to meet highly specialized needs of the host. Additionally, endogenous and exogenous<br />

mediators of immune and inflammatory response regulate specialized functions at the surface of the<br />

endothelium. The endothelium can be considered an effective mediator of organ homeostasis. 10,21<br />

In many ways, the vascular networks found in solid tumors poorly resemble the more regular,<br />

well-defined vascular structure observed in disease-free tissues. Tumors, for example, have a highly<br />

chaotic arrangement of vessels compared to vessels in normal tissues. Tumor vessels also have an<br />

over abundance of anionic phospholipids in addition to a number of other negatively charged<br />

functional groups. 22–25 In view of the negatively charged molecules, glycosaminoglycans carry<br />

out important functions in the metastatic disease process and, much like phospholipids, can serve as<br />

useful targets of cationic drug carrier molecules. Evaluation of altered proteoglycan expression in<br />

human breast tissue revealed a total proteoglycan content that was significantly increased in<br />

comparison to that in healthy tissues. 26 Proteoglycans isolated from malignant breast tissue have<br />

been shown to stimulate endothelial cell proliferation, and the total glycoprotein content in tumors<br />

is produced <strong>by</strong> many cell types, including cancer and tumor endothelial cells alike.<br />

The vascular networks of tumors have an increased permeability for macromolecules and a<br />

higher proliferation rate of endothelial cells compared to vessels in quiescent tissues. 27,28 An<br />

estimated 30- to 40-fold increase in the growth rate of endothelial cells lining vessels in tumors<br />

compared to that in normal tissues has been demonstrated. 27<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


616<br />

Nanotechnology for Cancer Therapy<br />

Direct access to intravenously administered agents, rapid proliferation rate of endothelial cells,<br />

and over-expression of negatively charged functional groups along vessels are potentially exploitable<br />

features of tumors. Endothelium-specific cationic liposomes made more specific through<br />

conjugation of antibodies can offer tremendous therapeutic gains.<br />

30.3 ARGUMENTS FOR VASCULAR TARGETING APPROACH OVER<br />

INTERSTITIAL DRUG TARGETING<br />

Physiological barriers that prevent effective interstitial delivery and drug transport 29 will likely pose<br />

little threat to vascular targeting strategies given that blood vessels have better access to intravenously<br />

administered agents. Damage to only a few tumor vessels can result in the death of literally thousands<br />

of malignant cells. 30 Limited drug diffusion to hard-to-reach interstitial areas in tumors will ultimately<br />

spare clusters of well-differentiated cancer cells from treatment, but direct targeting of cancer cells is<br />

not as important for vascular targeting strategies. As a therapeutic approach, vascular targeting is<br />

unaffected <strong>by</strong> convection and diffusion, and it does not require the presence of highly perfused vessels<br />

in all tumor regions to be effective. It can be argued that irregular, slow blood flow velocities play a<br />

critical role in facilitating interaction of drugs with endothelial cells. 31 Additionally, poorly perfused<br />

tumor vessels permit limited access of drugs to cancer cells. Additionally, endothelial cells are nonmalignant<br />

and are far less likely than cancer cells to acquire the mutations that safeguard them from<br />

treatment. Finally, should a drug intended to target functional tumor vessels gain direct access to a<br />

population of cancer cells in the process, additional gains for therapy could result. The potential for<br />

vascular targeting substances to come in direct contact with cancer cells can be observed in mosaic<br />

vessels where tumor cells directly interact with the lumen of vessels and flowing blood and assist in<br />

the spread of metastatic disease. 32 This is apparent with large tumor vascular pore openings where<br />

drug agents have considerably greater access to malignant cell populations. 11<br />

Proof-of-principal: A significant amount of experimental evidence can now be used to<br />

support vascular targeting as a targeted approach against cancer. To illustrate, tumor-bearing<br />

mice treated with doxorubicin–RGD-4C conjugate (tumor-homing peptide), outlived control<br />

mice that died from widespread disease. 33 Mice treated every 3 weeks for a total of 12 weeks<br />

with doxorubicin-RGD-4C lived 6 months longer compared to doxorubicin-treated mice. 33 An<br />

in vitro study showed success with immunoliposome-targeted delivery to ICAM-1-expressing<br />

cells in the bronchial epithelium and endothelium. 34 In this study, the extent of ICAM-1<br />

expression was suggested to influence the degree of immunoliposome binding to endothelial<br />

cells. The immunotargeting of liposomes to activated vascular endothelial cells of the cardiovascular<br />

system has also been investigated. 35 In this study, murine antibody mAB H18/7 was<br />

conjugated to doxorubicin-loaded liposomes, and the extracellular binding domain of E-selectin<br />

was targeted to a significant extent over nonactivated human endothelial cells. 35 Thorpe and<br />

colleagues characterized TEC-11 antibody and have demonstrated that the anti-class-ricin<br />

A-chain immunotoxin produced widespread infarction and tumor regression in treated mice<br />

compared to untreated controls. 36 While these groups and others work to demonstrate proofof-principal,<br />

other laboratories have had success with identifying accessible tumor vessel<br />

targets. One example of a potentially useful clinical target was studied <strong>by</strong> Retting and<br />

colleagues. They confirmed that the tumor endothelial antigen endosialin (cell glycoprotein)<br />

is expressed in tumors and is not expressed in normal healthy tissues. 37 This early finding<br />

suggests that, under optimized conditions, the glycoprotein could be used to launch a sitespecific,<br />

immunological attack on tumor vessels. Multiple lines of evidence from the literature<br />

support vascular targeting as an attractive therapeutic approach against cancer disease.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 617<br />

30.4 INFLUENCE OF CATIONIC CHARGE ON VASCULAR TARGETING<br />

IN HEALTH AND IN DISEASE<br />

To optimally target drugs to the endothelium, a basic working knowledge of why select carriers<br />

accumulate at this site is essential. Extraordinary insights into normal vessel function and pathogenesis<br />

now provide some interesting clues. For example, the vessel lumen of capillaries is<br />

predominately lined with anionic sites. 22–25 These regions are important in the development and<br />

maintenance of normal and abnormal tissues where the latter is observed in inflammation, thrombosis,<br />

and in tumors. 24,25,38 Anionic sites (owing to layers of glycoproteins) mediate microvascular<br />

permeability and trans-endothelial cell function. 3,6,23–25,30,38 Endothelial cell membrane-associated<br />

proteoglycans regulate delivery and uptake of macromolecules. This has been demonstrated in<br />

sulfated proteoglycan deficient HeLa cells where cation-mediated delivery of plasmid DNA<br />

resulted in a 69% reduction in luciferase expression. 39 The distribution of negatively charged<br />

sites is non-uniform, 25,31 varying among vessel types (i.e., arteries, capillaries, or capillary<br />

venules) and between young and relatively old mice. 4 Although negatively charged components<br />

of mammalian cells are sialoglycoproteins and proteoglycans, some vascular domains (potentially<br />

involved in endocytosis and/or trancytosis) are devoid of these functional groups. These specialized<br />

areas interact to a lesser extent with cationized molecules, and it has been suggested that they<br />

represent regions of anionic molecular transport across capillary networks. 25 The absence of<br />

accessible anionic sites can be observed on the vascular surface, specifically on the surface of<br />

plasmalemmal vesicles where they readily assist in the transport of circulating macromolecules. 4,6<br />

Experiments in solid tumors support the notion that success with nanoparticle-assisted delivery<br />

of drugs to tumor vessels does not require that drug carrier molecules possess a cationic charge for<br />

effective delivery and transport. In fact, when anionic sites were neutralized with the polycation<br />

protamine, the blood–brain barrier (BBB) became significantly less guarded against the transport of<br />

albumin and inulin; under normal conditions, neither can pass the BBB. 40 Additionally, whereas<br />

electron microscopic studies revealed that positively charged electron-dense label cationic ferritin<br />

(CF) did not label plasmalemmal vesicles of the mouse endothelium, native anionic and neutralized<br />

ferritin molecules had considerably greater access to these sites. 4 Another study showed that<br />

following a cerebral insult of a rat brain tumor model, the vasculature was more permeable to<br />

horse radish peroxidase, but more importantly, showed a fewer number of negatively charged sites.<br />

The mechanism(s) <strong>by</strong> which anionic surface charge is altered in disease is not completely understood,<br />

40–42 but studies continue to support structural and morphological differentiation of the<br />

microvascular endothelium in disease tissues that undoubtedly influences overall vascular<br />

targeting efficiency.<br />

The endothelium is structurally and functionally unique, and several studies confirm regionspecific,<br />

tumor vascular sensitivity to effects of molecular charge. 3,6,18,25,40,43,44 The fact that the<br />

endothelium can bind anionic, neutral, and net positively charged molecules supports these observations.<br />

The strategies that best exploit the expression of negatively charged functional groups in<br />

combination with additional tumor vascular features will achieve the greatest therapeutic success.<br />

30.5 FORMULATION DEVELOPMENT: ALTERNATIVES TO ROUTINE DRUG<br />

DESIGN AND SYNTHESIS<br />

The development of novel therapeutics is an expensive and labor-intensive process. There is also no<br />

guarantee that a more useful therapeutic agent will result from the time and money invested. The use of<br />

combinatorial chemistry, high throughput screening, and drug synthesis has not met all clinical expectations.<br />

As a result, serious issues still remain with existing treatment approaches and with a search for a<br />

magic bullet. 45 For this reason, strategies that improve tumor endothelial site-specific delivery are<br />

being developed in parallel with efforts involving routine drug design and synthesis.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


618<br />

In many instances, the use of nanotechnology represents a wise alternative to routine drug<br />

synthesis. This technology creates new opportunities to improve drug access to primary targets and<br />

to exploit additional tumor targets as well. Long circulating liposomes, for example, increase<br />

deposition of drugs in the perivascular space of tumors. 46,47 When the optimal lipid ratios have<br />

been considered, the vasculature may be targeted more selectively compared to free drug.<br />

On the basis of the experimental evidence, a non-toxic phospholipid-based delivery vehicle is a<br />

rational choice for targeting many therapeutic agents to tumors. Liposomes are well suited for this<br />

purpose. These microparticulates spontaneously form and are generally easier to prepare compared<br />

to viral mediated systems.<br />

Liposomes are relatively non-immunogenic; their large-scale manufacturing does not depend<br />

on the culturing of living cells, and they are preferred for a variety of practical considerations as<br />

well. Liposomes have been shown to increase systemic circulation, reduce drug side effects,<br />

enhance cellular uptake, and increase the duration of drug exposure and tumor response to<br />

therapy. 16–18 Moreover, the physicochemical properties of liposomes can be manipulated to<br />

address physiological considerations such as the effects of size, molecular charge, membrane<br />

rigidity, and stability. 48–53<br />

Developing a target specific, long circulating, stable drug carrier is only half the battle. Ideally,<br />

liposomes should also be triggered to release drugs under tumor sensitive conditions. The effect of<br />

molecular charge, over-expression of vascular growth factors, sensitivity to pH, and relatively large<br />

vascular pore openings are a few physiological features currently used to distinguish tumors from<br />

healthy organ tissues. 12,25,31,38,46,53–57<br />

30.6 INTERACTION OF CATIONIC LIPOSOMES WITH HUMAN<br />

ENDOTHELIAL CELLS<br />

Nanotechnology for Cancer Therapy<br />

One of the first synthetic cationic lipids, developed <strong>by</strong> Felgner and colleagues in 1987, 58,59 was<br />

initially used to improve existing DNA delivery methods. A 1:1 w/w mixture of the new cationic<br />

lipid DOTMA (N-[1-(2,3-dioleyloxy) propyl]-N,N,N-trimethylammonium chloride) was combined<br />

with the helper lipid DOPE (dioleoylphosphatidylethanolamine). This formulation (akaw-<br />

Lipofectin) was found to be 5–100 times more effective than DNA transfection performed with<br />

other in vitro methods. 59 A typical eukaryotic cell membrane has a highly negative surface charge<br />

density because of the multitude of membrane-bound carbohydrate groups oriented on its surface.<br />

Endothelial cells can be targeted on the basis of this surface charge density, and several<br />

mechanisms of action have already been proposed for successful delivery of DNA into cells <strong>by</strong><br />

positively charged liposomes. 60–63<br />

Studies involving the transfection of human endothelial cells suggest that methods be tailored<br />

to specifications of endothelial cells, not more traditional cell types. High transfection efficiency<br />

with the use of cationic liposomes is possible when expression is under the control of a strong<br />

promoter. 64 Endothelial cells, like most mammalian cells, depend on an optimal DNA/cationic<br />

liposome ratio for high efficiency of endothelial cell trafficking and expression of foreign genes.<br />

Moreover, compared to other mammalian cell lines, primary endothelial cells are generally more<br />

sensitive to constituents of growth medium and passage number. 65 No specific cationic lipid, to<br />

date, is accepted for use with all mammalian cell types. Cationic liposomes may represent a suitable<br />

alternative to non-viral delivery systems; however, transfection of human endothelial cells <strong>by</strong> this<br />

approach should be optimized to produce the desired therapeutic effect. 64<br />

Various experimental and clinical applications for cationic lipids have been reported, ranging<br />

from optimizing methods for enhanced gene transfer in vitro to treating clinical abnormalities<br />

contributing to neoplastic transformations in man. 66 Using cationic lipids to regulate endothelial<br />

cell function in vitro or in treatment of disease in highly developed organ systems has been<br />

explored. To illustrate, the application of several cell line types have been investigated<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 619<br />

FIGURE 30.1 (See color insert following page 522.) Cationic liposomes are taken up <strong>by</strong> human endothelial<br />

cells. The figure shows a DIC and fluorescence merged image of PCLs associated with human dermal<br />

endothelial cells. Perinuclear localization is observed. Captured image is representative of PCLs taken up<br />

<strong>by</strong> other primary endothelial cells derived from various organ tissues. Cells were seeded at 5!10 5 cells/well on<br />

cover slips in six well plates. Images were captured 24 h after exposure of cells to PCLs.<br />

(Figure 30.1). Cationic liposome-mediated transfection has been carried out under a broad set of<br />

experimental conditions, involving the use of numerous in vitro models of the endothelium in<br />

human organ tissues. These in vitro models of the intravascular compartment include the eye, 67<br />

lung, 68,69 heart, 61,70 kidney, 71 and vascular smooth muscle. 72,73<br />

Site-specific delivery to endothelial cells can be accomplished <strong>by</strong> coating cationic lipsoemes<br />

with peptides and monoclonal antibodies. 65,74 Furthermore, these positively charged vehicles<br />

enhanced adeno-associated virus (AAV) uptake <strong>by</strong> human endothelial and smooth muscle cells.<br />

The authors report a highly efficient way to introduce foreign DNA into cells given that modified<br />

cationic liposomes (adenosomes) enhanced vector expression <strong>by</strong> a factor of 10 over adenoviral<br />

proteins delivered without liposomes. 75<br />

Although antibodies have been used to assist gene transfer strategies, cationic liposomes can be<br />

optimized to deliver genes to endothelial cells without the use of antibodies. For instance, when<br />

cultured bovine corneal endothelial cells (BCECs) were exposed to pUT651 (a plasmid) alone,<br />

significant expression was observed and without associated toxicities. 67 In HUVEC cells, reporter<br />

gene expression was evident in the absence of select antibodies. In this study, transfection efficiency<br />

varied as a function of cell passage number, and a linear reduction in luciferase expression was<br />

observed with successive passage numbers. 65 Vascular smooth muscle cells were shown to vary in<br />

degree of sensitivity to transfection reagents mediated <strong>by</strong> cationic lipids, reporting a preference for<br />

restenotic lesions compared to normal internal thoracic arteries and atherosclerotic plaques. 73<br />

30.7 STEALTH (PEGYLATED) LIPOSOME TECHNOLOGY: A BRIEF<br />

COMMUNICATION<br />

Rapid elimination of cationic liposomes from circulation <strong>by</strong> opsonins, lipoproteins, and cell-surface<br />

receptors pose an additional threat to effective delivery of drugs to solid tumors. 45,76–80 This<br />

particular barrier is different from physiological barriers as it results not from tumors or the<br />

influences from the surrounding tissue environment, but it occurs when conventional (relatively<br />

short circulating) liposomal therapeutics come in direct contact with circulating proteins in blood.<br />

Stealth liposome technology is one approach used to limit the interaction of conventional<br />

liposomes with proteins in blood. The technology involves the coupling of high molecular<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


620<br />

weight polymers to the liposome surface. 47,50 Surface-bound polymers [i.e., poly(ethylene glycol)<br />

(PEG)] provide adequate protection of liposomes from protein molecules that can unfavorably alter<br />

their tissue distribution profile. The specific mechanisms underlying delivery mediated <strong>by</strong> these<br />

polymers are unclear, but when PEG is included as a component of liposome preparations, a<br />

physical zone is formed around the liposome perimeter. This zone of exclusion reduces liposome–protein<br />

interactions as a result of long (PEG) polymer chain constrictions. 47,50,80<br />

From a therapeutic point-of-view, in addition to favorable kinetics, PEGylated liposomes are<br />

usually more active against solid tumors. 47,79 Several lines of evidence support the findings that<br />

PEGylated liposomes are generally more stable in blood with only minor leakage of their contents<br />

over time compared to conventional liposomal therapeutics. 78<br />

PEG can also be used to prepare liposomes with significantly higher drug retention under<br />

normal physiological pH, or it can be designed to trigger drug release in a low pH environment.<br />

In addition to tumor uptake, other organs of relatively high accumulation are the liver, kidney, skin,<br />

the female reproductive, and gastrointestinal tract. 77,78<br />

The extent of modification and molecular mechanics of Stealth liposome technology and<br />

exactly how PEG delays reticuloendothelial system (RES) entrapment is not completely understood.<br />

However, its benefit(s) in systemic circulation and that the fraction of PEG included in<br />

Stealth liposome preparations should reflect preclinical and clinical goals are understood.<br />

Stealth technology might well represent one of the most significant advances in the field of<br />

liposome nanotechnology. PEG has provided researchers with an enormous opportunity to exploit<br />

the large openings of tumor vessels <strong>by</strong> enhancing delivery of therapeutic substances through them<br />

as a result of the advantage of prolong circulation. In a study performed <strong>by</strong> Yuan et al., Stealth<br />

liposomes accumulated in the perivascular space in tumors 48 h post-administration. 46 This is the<br />

region close to tumor vessels but not directly associated with them. The technology is, however,<br />

not sufficient to promote vascular-specific delivery; targeting specific physiological feature of<br />

vessels (i.e., proteoglycans, endothelial receptors, etc.) in addition to prolonged circulation is<br />

also required.<br />

30.8 PEGYLATED CATIONIC LIPOSOMES: ARE THEY SUITED<br />

FOR TARGETING TUMOR VESSELS?<br />

Nanotechnology for Cancer Therapy<br />

The notion of targeting accessible anionic sites along tumor vessels combined with technologies<br />

that limit interactions of nanodelivery systems with biological molecules in vivo is promising.<br />

PEGylated cationic liposomes (PCLs) have been shown to improve not only oligonucleotide<br />

(ODN) loading and delivery, but also the apparent drug solubility profile of various therapeutic<br />

molecules as well. In vitro studies conducted with human plasma show that PCLs do not form large<br />

molecular aggregates in blood 81,82 or in buffered solutions. 31 These carriers retain most of their<br />

physicochemical properties, 31,75,81 and they prolong circulation half-life even in the presence of a<br />

net positively charged surface.<br />

Protecting cationic liposomes from serum nucleases enhanced immunostimulatory activity of<br />

CpG oligonucleotides motifs compared to free ODN in in vitro and in mouse models. 82 In another<br />

study, the anti-HER-2 F (ab 0 ) fragment was coupled to the distal end of PEG chains of cationic<br />

liposomes; cellular uptake and nuclear localization increased considerably in cultured SK-BR-3<br />

cells (a human breast cancer cell line) shown to over-expresses HER-2 oncoprotein. 81<br />

Intravital microscopy has been used to investigate the interactions of PCLs with tumor vessels<br />

in vivo. 31 When PCLs were intravenously injected into tumor bearing (LS174T-human colon<br />

adenocarcinoma) mice, analysis of spatial distribution of liposomes 24 h post-injection revealed<br />

significant vascular targeting in tumors (w27.5% coverage) compared to vessels in normal tissues<br />

(w3.6% coverage). 31 Moreover, the extent of vascular areas targeted did not vary as a function of<br />

tumor type or anatomical location. 31 Several lines of evidence suggest that PEG coating does not<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 621<br />

inhibit cationic liposomes from binding to negatively charged surfaces. In addition to experimental<br />

data supporting this hypothesis, the values of zeta potential (electric potential across a double<br />

membrane surface) confirm that a sufficient electrostatic membrane potential still exists in the<br />

presence of PEG. 31,52,83<br />

Vascular targeting can be influenced on the basis of charge content of drug carrier molecules;<br />

associations with tumor vessels favor cationic liposomes possessing relatively high cationic charge<br />

content. 31,38,63 The abilities to circulate for extended periods in blood and associate with tumor<br />

vessels are highly desirable features of an efficient therapeutic approach. For this reason, both the<br />

inclusion of PEG and cationic charge content in cationic liposomes are currently being investigated<br />

in combination. There are clinical circumstances when relatively long circulating drug carrier<br />

molecules (O24 h) accumulating along tumor vessels would be more effective than carriers that<br />

are rapidly eliminated from circulation (!0.5 h). For instance, many of the popular chemotherapeutic<br />

agents are dependent on phase of the cell cycle; therefore, rapidly dividing tumor endothelial<br />

cells could be targeted more effectively with carriers that maintained adequate levels of drug in<br />

blood. Such a delivery system would serve less value for agents that can exert drug effects outside<br />

the influence of the cell cycle (i.e., membrane targeting agents).<br />

30.9 CATIONIC LIPOSOMES IMPROVE ANTI-CANCER THERAPY<br />

The use of cationic liposomes to deliver experimental therapeutics to solid tumors is a promising<br />

alternative to interstitial targeting approaches. However, formulations should be optimized and prescreened<br />

in preclinical models to establish clinical relevance. Given the fact that the architecture of<br />

tumor vessels varies significantly in terms of microvascular type, density, vascular surface area, and<br />

angiogenic potential, it is understandable why the development of a single therapeutic approach<br />

against cancer has yet to be achieved.<br />

When developing cationic liposomal therapeutics, it will be helpful to note that each solid<br />

tumor possesses distinctly different organization and specialized structural features. The highly<br />

unpredictable nature of tumors further complicates the task of pairing formulations to specific<br />

disease states. A common feature among all cationic liposome preparations used in drug delivery<br />

is that they must be optimized for site-specific delivery to tumor endothelia and that their intracellular<br />

fate should correlate with the drug’s mechanism of action (i.e., cationic liposomes should<br />

enhance delivery of DNA targeting agents to DNA, and not to irrelevant intracellular targets). This<br />

process must be confirmed.<br />

Cationic liposomes can retain drug agents at the tumor vascular site (Figure 30.2), and facilitate<br />

interaction of liposomes with subcellular targets prior to releasing their payload. They can also be<br />

used to target non-intracellular targets as well as cell-membrane bound molecules other than<br />

proteoglycans. This is promising given that many anti-angiogenic agents have been confirmed to<br />

exert their action <strong>by</strong> each mechanism. For instance, SU5416, an inhibitor of tyrosine kinase activity<br />

of vascular endothelial growth factor (VEGF), requires direct access to a specific endothelial cell<br />

membrane-associated receptor (Flk-1/KDR) in order to suppress neovascular growth of tumors. 84,85<br />

Although SU5416 is suggested to exert long-lasting effects on VEGF phosphorylation and function,<br />

cationic liposome-assisted drug delivery could enhance interactions with specific endothelial<br />

cell targets.<br />

Effective anti-angiogenic therapy requires the continuous presence of drugs in circulation. 85,86<br />

The inclusion of PEG in cationic liposome preparations can extend circulation half-life of SU5416<br />

compared to SU5416 alone; the goal here is to enhance the duration of drug (SU5416) exposure<br />

with tumor target.<br />

Most anti-angiogenic agents suppress neovascularization <strong>by</strong> blocking activators of angiogenesis<br />

and endothelial cell-specific signals. There exist clinical situations where targeting nontumoral<br />

tissues might be equally beneficial. Although the majority of PCLs are recovered <strong>by</strong> the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


622<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 30.2 (See color insert following page 522.) Targeting tumor vessels with cationic liposomes. The<br />

image shows vascular targeting of LS174T (human colon adenocarcinoma) with cationic liposomes and was<br />

captured <strong>by</strong> two-photon microscopy 250 mm deep within the tumor. The intrinsic fluorescent property of<br />

doxorubicin was used to probe the location of the drug within the tumor. When doxorubicin was loaded in<br />

PCLs and administered via tail vein in tumor bearing mice, the drug was shown to be delivered to tumor<br />

vessels. Green represents tumor vessels, and orange dots represent doxorubicin. Mag. BarZ50 mm.<br />

liver post-intraveneous 31 and intrasplenic 87 administration, once these carrier molecules have been<br />

loaded with chemotherapeutic agents, they can diminish primary tumor growth in this anatomical<br />

location without significantly altering its primary function. To illustrate, poly(I):poly(C)–cationic<br />

liposome complex was used to treat metastatic (pancreatic and colorectal) carcinoma cell growth in<br />

the liver, and the result was significant delay in tumor growth compared to the administration of<br />

poly(I):poly(C), and adriamycin alone. 87<br />

In a study using PCLs to deliver the chemotherapeutic agent cisplatin (CDDP), CDDP-loaded<br />

PCLs suppressed primary tumor growth and distant liver metastasis as a result of tumor and liverselective<br />

uptake. 88 The authors note that the high chondroitin sulfate content of cancer cells was the<br />

primary target of PCLs. In addition to this mechanism, another possibility is that drug loaded-PCLs<br />

deliver drugs like CDDP to tumor vessels and that cancer cells are not the primary, but the<br />

secondary, target of drug-loaded PCLs. Depending on the overall clinical objectives, liver accumulation<br />

could result in some unwanted complications; for wide-spread general applicability, high<br />

tumor selectivity at the expense of the liver is desired. To minimize uptake of cationic carriers at<br />

this anatomical site, additional specialized features of tumor vessels may have to be taken<br />

into account.<br />

In addition to targeting tumor vascular targets on the basis of negatively charged functional<br />

groups, cationic drug carriers can also be used to enhance the interaction of anti-angiogenic agents<br />

with integrins. Integrins are involved in cell invasion and metastasis and in signaling processes that<br />

regulate the actions of each. 89,90 Some integrins are preferentially expressed in angiogenic endothelium<br />

like integrin avb3. 89 Researchers have since introduced a novel cationic delivery vehicle in<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 623<br />

order to exploit this endothelial cell receptor. 91 In as study carried out <strong>by</strong> Hood and colleagues, a<br />

cationic lipid based nanoparticle (NP) was coupled to avb3-targeting ligand and intravenously<br />

injected in tumor bearing mice; the complex (avb3–NP) selectively targeted the vasculature and<br />

induced tumor regression. 91 Furthermore, sustained regression was achieved without significant<br />

accumulation in, and damage to, the liver. This finding is encouraging. It supports the use of<br />

positively charged drug carriers with highly specialized molecules to improve vascular targeting<br />

at all stages of tumor development.<br />

The knowledge that cationic liposomes deliver drugs to solid tumors as well as the degree to<br />

which each tumor compartment is affected <strong>by</strong> treatment is essential to future advances in nanotechnology.<br />

Unfortunately, the overwhelming majority of published works document overall tumor<br />

response, animal survival, and drug effects involving the interstitial matrix, yet hidden<br />

mechanism(s) underlying tumor vascular damage in connection with treatment is often not a part<br />

of routine investigations.<br />

No intravenously administered substances (drug alone or nanoparticle-assisted) can gain access<br />

to the interstitial matrix without first passing through the vascular compartment, so drug effects to<br />

this tumor compartment must be explored. Cationic liposomes possessing a high cationic lipid<br />

content (w50 mol%) target the endothelium several-fold over the interstitial environment. 31,38<br />

Growth inhibition and animal survival studies have proven to be extremely helpful in establishing<br />

safety conditions for drug administration and predicting drug efficacy in humans. Nonetheless, the<br />

specific mechanisms <strong>by</strong> which cationic liposomes target vessels and alter tumor growth kinetics<br />

must be investigated. Studies should probe interstitial and vascular compartments without<br />

experimental bias.<br />

30.10 VALIDATION AND LONG-TERM TRACKING OF TUMOR RESPONSE<br />

TO THERAPY<br />

It has been established that there are unique molecular and physiological features of tumors that can<br />

be used to develop relatively disease-specific treatments. The expression of tumor-specific genes<br />

has been used to monitor progression of disease in response to these treatments and others. For<br />

instance, VEGF is a potent angiogenic cytokine that is critically linked to angiogenesis; it is so vital<br />

to this process that inhibitors are rapidly being developed to suppress VEGF’s tumor angiogenesis<br />

stimulatory activity. 84,92–94 Normally, when select inhibitors of VEGF bind Flt-1 (VEGF target<br />

site) on endothelial cells, proliferation is inhibited among other critical steps involved in tumor<br />

angiogenesis. 20,30,84 VEGF expression is a highly regulated process in healthy tissue; elevated<br />

expression is observed only under special circumstances, but constitutive expression is observed<br />

in tumors that aggressively seek to recruit a new blood supply. Given the fact that the process of<br />

developing new blood vessels is fundamentally important to maintaining an aggressive tumor<br />

phenotype, VEGF and a variety of other growth factors are often used as molecular and<br />

pharmacological endpoints.<br />

Site-specific delivery to tumor vessels and angiogenesis has demonstrated promise in preclinical<br />

models, warranting follow-up investigations in clinical settings. Evaluating tumor volume,<br />

animal survival and vascular density are useful indicators of tumor response. However, monitoring<br />

gene expression in response to therapy can provide deeper insights into the specific mechanisms<br />

underlying the disease process.<br />

30.11 CONCLUDING REMARKS<br />

Systemic administration of drugs is a frequently applied strategy used to control local tumor growth<br />

and metastasis. The issue of targeting these agents more specifically to tumors while reducing<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


624<br />

uptake <strong>by</strong> healthy organ tissues is a formidable challenge. Cationic liposomes have been used to<br />

mediate efficient transfection of a variety of different mammalian cell types. The loading of cationic<br />

liposomes with chemotherapeutic agents has even greater potential. The relatively new class of<br />

liposomes possesses high cationic charge content (and membrane surface charge potential) similar<br />

to those used for transfection studies, but they deliver their drug payload to tumor vessels. The<br />

development of new synthetic cationic lipids like N-[1-(2,3-dioleyloxy)propyl]-N, N, N-trimethylammonium<br />

chloride(DOTMA) and N-[1-(2,3-dioleyloxy)propyl]-N, N, N-trimethylammonium<br />

chloride(DOTAP) has helped make this field a reality.<br />

A deeper understanding of how tumors develop relative to the expression of vascular receptors<br />

and other targets will create new opportunities for nanotechnologists. These opportunities extend to<br />

include more effective treatments at various stages of tumor development and planned attacks<br />

against cancer.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Jain, R. K., Vascular and interstitial barriers to delivery of therapeutic agents in tumors, Cancer<br />

Metastasis Rev., 9, 253–266, 1990.<br />

2. Simionescu, M., Simionescu, N., and Palade, G. E., Morphometric data on the endothelium of blood<br />

capillaries, J. Cell Biol., 60, 128–152, 1974.<br />

3. Simionescu, M., Simionescu, N., and Palade, G. E., Structural basis of permeability in sequential<br />

segments of the microvasculature. I. Bipolar microvascular fields in the diaphragm, Microvasc. Res.,<br />

15, 1–16, 1978.<br />

4. Simionescu, M., Simionescu, N., Santoro, F., and Palade, G. E., Differentiated microdomains of the<br />

luminal plasmalemma of murine muscle capillaries: Segmental variations in young and old animals,<br />

J. Cell Biol., 100, 1396–1407, 1985.<br />

5. Ghinea, N. and Simionescu, N., Anionized and cationized hemeundecapeptides as probes for cell<br />

surface charge and permeability studies: Differentiated labeling of endothelial plasmalemmal<br />

vesicles, J. Cell Biol., 100, 606–612, 1985.<br />

6. Simionescu, M., Simionescu, N., and Palade, G. E., Differentiated microdomains on the luminal<br />

surface of the capillary endothelium. I. partial characterization of their anionic sites, J. Cell Biol.,<br />

90, 614–621, 1981.<br />

7. Movat, H. Z., The role of histamine and other mediators in microvascular changes in acute inflammation,<br />

Can. J. Physiol. Pharmacol., 65, 451–457, 1987.<br />

8. Mylecharane, E. J., Mechanisms involved in serotonin-induced vasodilation, Blood Vessels, 27,<br />

116–126, 1990.<br />

9. Vercellotti, G. M. and Tolins, J. P., Endothelial activation and the kidney: Vasomediator modulation<br />

and antioxidant strategies, Am. J. Kidney Dis., 21, 331–343, 1993.<br />

10. van Hinsbergh, V. W., The endothelium: Vascular control of haemostasis, Atherosclerosis, 95,<br />

198–201, 2001.<br />

11. Hobbs, S. K. et al., Regulation of transport pathways in tumor vessels: Role of tumor type and<br />

microenvironment, Proc. Natl Acad. Sci. USA, 95, 4607–4612, 1998.<br />

12. Yuan, F. et al., Vascular permeability in a human tumor xenograft: Molecular size dependence and<br />

cutoff size, Cancer Res., 55, 3752–3756, 1995.<br />

13. Hudson, B. G., Reeders, S. T., and Trygvason, K., Type IV collagen: Structure, gene organization and<br />

role in human diseases, J. Biol. Chem., 268, 26033–26036, 1993.<br />

14. Gorog, P. and Pearson, J. D., Sialic acid moieties on surface glycoproteins protect endothelial cells<br />

from proteolytic damage, J. Pathol., 146, 205–212, 1985.<br />

15. Hallmann, R. et al., Expression of function of laminins in the embryonic and mature vasculature,<br />

Physiol. Rev., 85, 979–1000, 2005.<br />

16. Sund, M., Xie, L., and Kalluri, R., The contribution of vascular basement membranes and extracellular<br />

matrix to the mechanics of tumor angiogenesis, APMIS, 112, 450–462, 2004.<br />

17. Ruoslahti, E., Structure and biology of proteoglycans, Annu. Rev. Cell Biol., 4, 229–255, 1988.<br />

18. Ruoslahti, E., Specialization of tumor vasculature, Nat. Rev. Cancer, 2, 83–90, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 625<br />

19. Satoh, A., Toida, T., Yoshida, K., Kojima, K., and Matsumoto, I., New role of glycosaminoglycans on<br />

the plasma membrane proposed <strong>by</strong> their interaction with phosphatidylcholine, FEBS Lett., 477,<br />

249–252, 2000.<br />

20. Folkman, J. and Haudenschild, C., Angiogenesis <strong>by</strong> capillary endothelial cells in culture, Trans.<br />

Ophthalmol. Soc. UK, 100, 346–353, 1980.<br />

21. Michiels, C., Arnould, T., and Remacle, J., Endothelial cell responses to hypoxia: Initiation of a<br />

cascade of cellular interactions, Biochim. Biophys. Acta, 1497, 1–10, 2000.<br />

22. Charonis, A. S. and Wissig, S. L., Anionic sites in basement membranes. Differences in their electrostatic<br />

properties in continuous and fenestrated capillaries, Microvasc. Res., 25, 265–285, 1983.<br />

23. Roberts, W. G. and Palade, G. E., Neovasculature induced <strong>by</strong> vascular endothelial growth factor is<br />

fenestrated, Cancer Res., 57, 765–772, 1997.<br />

24. Ran, S., Downes, A., and Thorpe, P. E., Increased exposure of anionic phospholipids on the surface of<br />

tumor blood vessels, Cancer Res., 62, 6132–6140, 2002.<br />

25. Vincent, S., DePace, D., and Finkelstein, S., Distribution of anionic sites on the capillary endothelium<br />

in an experimental brain tumor model, Microcirc. Endothleium Lymphatics, 45–67, 1988.<br />

26. Vijayagopal, P., Figueoroa, J. E., and Levine, E. A., Altered composition and increased endothelial<br />

cell proliferative activity of proteoglycans isolated from breast carcinoma, J. Surg. Oncol., 68,<br />

250–254, 1998.<br />

27. Denekamp, J. and Hobson, B., Endothelial-cell proliferation in experimental tumours, Br. J. Cancer,<br />

46, 711–720, 1982.<br />

28. Denekamp, J., Vasculature as a target for tumor therapy, Prog. Appl. Microcirc., 4, 28–38, 1984.<br />

29. Jain, R. K., The next frontier of molecular medicine: Delivery of therapeutics, Nat. Med., 4, 655–657,<br />

1998.<br />

30. Folkman, J., Tumor angiogenesis: Therapeutic implications, N. Engl. J. Med., 285, 1182–1186, 1971.<br />

31. Campbell, R. B. et al., Cationic charge determines the distribution of liposomes between the vascular<br />

and extravascular compartments of tumors, Cancer Res., 62, 6831–6836, 2002.<br />

32. Chang, Y. S. et al., Mosaic blood vessels in tumors: Frequency of cancer cells in contact with flowing<br />

blood, Proc. Natl Acad. Sci. USA, 97, 14608–14613, 2000.<br />

33. Arap, W., Pasqualini, R., and Ruoslahti, E., Cancer treatment <strong>by</strong> targeted drug delivery to tumor<br />

vasculature in a mouse model, Science, 279, 377–380, 1998.<br />

34. Bloemen, P. G. et al., Adhesion molecules: A new target for immunoliposome-mediated drug<br />

delivery, FEBS Lett., 357, 140–144, 1995.<br />

35. Spragg, D. D. et al., Immunotargeting of liposomes to activated vascular endothelial cells: A strategy<br />

for site-specific delivery in the cardiovascular system, Proc. Natl Acad. Sci. USA, 94, 8795–8800,<br />

1997.<br />

36. Thorpe, P. E. and Burrows, F. J., Antibody-directed targeting of the vasculature of solid tumors,<br />

Breast Cancer Res. Treat., 36, 237–251, 1995.<br />

37. Rettig, W. J. et al., Identification of endosialin, a cancer surface glycoprotein of vascular endothelial<br />

cells in human cancer, Proc. Natl Acad. Sci. USA, 89, 10832–10836, 1992.<br />

38. Thurston, G. et al., Cationic liposomes target angiogenic endothelial cells in tumors and chronic<br />

inflammation in mice, J. Clin. Investig., 101, 1401–1413, 1998.<br />

39. Mislick, K. A. and Baldeschwieler, J. D., Evidence for the role of proteoglycans in cation-mediated<br />

gene transfer, Proc. Natl Acad. Sci. USA, 93, 12349–12354, 1996.<br />

40. Hardebo, J. E. and Kahrstrom, J., Endothelial negative surface charge areas and blood–brain barrier<br />

function, Acta Physiol. Scand., 125, 495–499, 1985.<br />

41. Nag, S., Role of the endothelial cytoskeleton in blood–brain-barrier permeability to proteins, Acta<br />

Neuropathol. (Berl.), 90, 454–460, 1995.<br />

42. Nag, S., Blood–brain barrier permeability using tracers and immunohistochemistry, Methods Mol.<br />

Med., 89, 133–144, 2003.<br />

43. Renkin, E. M., Transport of macromolecules: Pores and other endothelial pathways, Appl. Physiol.,<br />

58, 315–325, 1985.<br />

44. Renkin, E. M., Multiple pathways of capillary permeability, Circ. Res., 41, 735–743, 1977.<br />

45. Gabizon, A. A., Stealth liposomes and tumor targeting: One step further in the quest for the magic<br />

bullet, Clin. Cancer Res., 7, 223–225, 2001.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


626<br />

Nanotechnology for Cancer Therapy<br />

46. Yuan, F. et al., Microvascular permeability and interstitial penetration of sterically stabilized (Stealth)<br />

liposomes in a human tumor xenograft, Cancer Res., 54, 3352–3356, 1994.<br />

47. Papahadjopoulos, D. et al., Sterically stabilized liposomes: Improvements in pharmacokinetics and<br />

anti-tumor efficacy, Proc. Natl Acad. Sci. USA, 88, 11460–11464, 1991.<br />

48. Campbell, R. B., Balasubramanian, S. V., and Straubinger, R. M., Physical properties of phospholipid–cationic<br />

lipid interactions: Influences on domain structure, liposome size and cellular uptake,<br />

Biochim. Biophys. Acta, 1512, 27–39, 2001.<br />

49. Campbell, R. B., Balasubramanian, S. V., and Straubinger, R. M., Influence of cationic lipids on the<br />

stability and membrane properties of paclitaxel-containing liposomes, J. Pharm. Sci., 90, 1091–1105,<br />

2001.<br />

50. Torchilin, V. P., Polymer-coated long circulating microparticulate pharmaceuticals, J. Microencapsul.,<br />

15, 1–19, 1998.<br />

51. Szoka, F. and Papahadjopoulos, D., Comparative properties and methods of preparation of lipid<br />

vesicles (liposomes), Annu. Rev. Biophys. Bioeng., 9, 467–508, 1980.<br />

52. Levchenko, T. S., Rammohan, R., Lukyanov, A. N., Whiteman, K. R., and Torchilin, V. P., Liposome<br />

clearance in mice: The effect of a separate and combined presence of surface charge and polymer<br />

coating, Int. J. Pharm., 240, 95–102, 2002.<br />

53. Straubinger, R. M., pH-Sensitive liposomes for delivery of macromolecules into cytoplasm of<br />

cultured cells, Methods Enzymol., 221, 361–376, 1993.<br />

54. Simberg, G., Weisman, S., Talmon, Y., and Barenholz, Y., DOTAP (and other cationic lipids):<br />

Chemistry, biophysics, and transfection, Crit. Rev. Ther. Drug Carrier Syst., 21, 257–317, 2004.<br />

55. Jain, R. K., Munn, L. L., and Fukumura, D., Dissecting tumour pathophysiology using intravital<br />

microscopy, Nat. Rev. Cancer, 2, 266–276, 2002.<br />

56. Hashizume, H. et al., Openings between defective endothelial cells explain tumor vessel leakiness,<br />

Am. J. Pathol., 156, 1363–1380, 2000.<br />

57. Fukumura, D., Yuan, F., Monsky, W. L., Chen, Y., and Jain, R. K., Effect of host microenvironment<br />

on the microcirculation of human adenocarcinoma, Am. J. Pathol., 151, 679–688, 1997.<br />

58. Felgne, P. et al., Lipofection: A highly efficient, lipid-mediated DNA-transfection procedure, Proc.<br />

Natl Acad. Sci. USA, 84, 7413–7417, 1987.<br />

59. Felgner, P. L. and Ringold, G. M., Cationic liposome-mediated transfection, Nature, 337, 387–388,<br />

1989.<br />

60. Sakurai, F. et al., Effect of DNA/liposome mixing ratio on the physicochemical characteristics,<br />

cellular uptake and intracellular trafficking of plasmid DNA/cationic liposome complexes and subsequent<br />

gene expression, J. Control. Release, 255–269, 2000.<br />

61. Dabbas, S., Kaushik, R., Dandamudi, S., and Campbell, R.B., Physiochemical evaluation of PEGmodified<br />

cationic liposomes (PCLs) and properties in vitro using human primary and immortalized<br />

microvascular endothelial cells. J. Pharm. Sci. Submitted.<br />

62. Dellian, M., Yuan, F., Trubetskoy, V. S., Torchilin, V. P., and Jain, R. K., Vascular permeability in a<br />

human tumor xenograft: Molecular charge dependence, Br. J. Cancer, 82, 1513–1518, 2000.<br />

63. McLean, J. W. et al., Organ-specific endothelial cell uptake of cationic liposome–DNA complexes in<br />

mice, Am. J. Physiol., 273, 387–404, 1997.<br />

64. Tanner, F., Carr, D. P., Nabel, G. J., and Nabel, E. G., Transfection of human endothelial cells,<br />

Cardiovasc. Res., 35, 522–528, 1997.<br />

65. Matsumura, J. S., Kim, R., Shively, V. P., MacDonald, R. C., and Pearce, W. H., Characterization of<br />

vascular gene transfer using a novel cationic lipid, J. Surg. Res., 85, 339–345, 1999.<br />

66. Nabel, G. J. et al., Direct gene transfer with DNA–liposome complexes in melanoma: Expression,<br />

biologic activity, and lack of toxicity in humans, Proc. Natl Acad. Sci. USA, 90, 11307–11311, 1993.<br />

67. Pleyer, U. et al., Efficiency and toxicity of liposome-mediated gene transfer to corneal endothelial<br />

cells, Exp. Eye Res., 73, 1–7, 2001.<br />

68. Uyechi, L. S., Gagne, L., Thurston, G., and Szoka, F. C. J., Mechanism of lipoplex gene delivery in<br />

mouse lung: Binding and internalization of fluorescent lipid and DNA components, Gene Ther., 8,<br />

828–836, 2001.<br />

69. Trubetskoy, V. S., Torchilin, V. P., Kennel, S., and Huang, L., Cationic liposomes enhance targeted<br />

delivery and expression of exogenous DNA mediated <strong>by</strong> N-terminal modified poly(L-lysine)–<br />

antibody conjugate in mouse lung endothelial cells, Biochim. Biophys. Acta, 1131, 311–313, 1992.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Positively-Charged Liposomes for Targeting Tumor Vasculature 627<br />

70. Abraham, N. G. et al., Transfection of the human heme oxygenase gene into rabbit coronary microvessel<br />

endothelial cells: Protective effect against heme hemoglobin toxicity, Proc. Natl Acad. Sci.<br />

USA, 92, 6798–6802, 1995.<br />

71. Nakatani, K. et al., Endothelial adhesion molecules in glomerular lesions: Association with their<br />

severity and diversity in lupus models, Kidney Int., 65, 1290–1300, 2004.<br />

72. Jeschke, M. G. et al., IGF-I gene transfer in thermally injured rats, Gene Ther., 6, 1015–1020, 1999.<br />

73. Pickering, J. G. et al., Liposome-mediated gene transfer into human vascular smooth muscle cells,<br />

Circulation, 89, 13–21, 1994.<br />

74. Tiukinhoy, S. D. et al., Development of echogenic, plasmid-incorporated, tissue-targeted cationic<br />

liposomes that can be used for directed gene delivery, Invest. Radiol., 35, 732–738, 2000.<br />

75. Hong, Z. et al., Enhanced adeno-associated virus vector expression <strong>by</strong> adenovirus protein–cationic<br />

liposome complex, Chin. Med. J., 108, 332–337, 1995.<br />

76. Torchilin, V. P., Liposomes as targetable drug carriers, CRC Crit. Rev. Ther. Drug Carrier Syst., 2,<br />

65–115, 1985.<br />

77. Ishida, T., Harashima, H., and Kiwada, H., Liposome clearance, Biosci. Reports, 22, 197–224, 2002.<br />

78. Harrington, K. J., Lewanski, C. R., and Stewart, J. S., Liposomes as vehicles for targeted therapy of<br />

cancer. Part 2. Clinical development, Clin. Oncol., 12, 16–24, 2000.<br />

79. Gabizon, A. and Paphadjopoulos, D., Liposome formulations with prolonged circulation time in blood<br />

and enhanced uptake in tumours, Proc. Natl Acad. Sci., 85, 6949–6953, 1985.<br />

80. Woodle, M. C. and Lasic, D. L., Sterically stabilized liposomes, Biochim. Biophys. Acta, 1113,<br />

171–199, 1992.<br />

81. Meyer, O. et al., Cationic liposomes coated with polyethylene glycol as carriers for oligonucleotides,<br />

J. Biol. Chem., 273, 15621–15627, 1998.<br />

82. Gursel, I., Gursel, M., Ishii, K. J., and Klinman, D. M., Sterically stabilized cationic liposomes<br />

improved the uptake and immunostimulatory activity of CpG oligonucleotides, J. Immunol., 167,<br />

3324–3328, 2001.<br />

83. Harigai, T. et al., Preferential binding of polyethylene glycol-coated liposomes containing a novel<br />

cationic lipid, TRX-20, to human subendothelial cells via chondroitin sulfate, Pharm. Res., 18,<br />

1284–1290, 2001.<br />

84. Strawn, L. M. and Shawver, L. K., Tyrosine kinase in disease: Overview of kinase inhibitors<br />

as therapeutic agents and current drugs in clinical trials, Expert Opin. Investig. Drugs, 7, 553–573,<br />

1998.<br />

85. Mendel, D. B. et al., The angiogenesis inhibitor SU5416 has long-lasting effects on vascular<br />

endothelial growth factor receptor phosphorylation and function, Clin. Cancer Res., 6, 4848–4858,<br />

2000.<br />

86. Brower, V., Tumor angiogenesis-new drugs on the block, Nat. Biotechnol., 17, 963–968, 1999.<br />

87. Hirabayashi, K., Yano, J., Takesue, H., Fujiwara, N., and Irimura, T., Inhibition of metastatic carcinoma<br />

cell growth in livers <strong>by</strong> poly(I):poly(C)/cationic liposome complex (LIC), Oncol. Res., 11,<br />

497–504, 1999.<br />

88. Lee, C. M. et al., Novel chondroitin sulfate-binding cationic liposomes loaded with cisplatin efficiently<br />

suppress the local growth and liver metastasis of tumor cells in vivo, Cancer Res., 62,<br />

4282–4288, 2002.<br />

89. Hood, J. D. and Cheresh, D. A., Role of integrins in cell invasion and migration, Nat. Rev. Cancer,2,<br />

91–100, 2002.<br />

90. Akalu, A., Cretu, A., and Brooks, P. C., Targeting integrins for control of tumour angiogenesis, Expert<br />

Opin. Investig. Drugs, 14, 1475–1486, 2005.<br />

91. Hood, J. D. et al., Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science, 296,<br />

2404–2407, 2002.<br />

92. Shinkarauk, S., Bayle, M., Lain, G., and Deleris, G., Vascular endothelial cell growth factor (VEGF),<br />

an emerging target for cancer therapy, Curr. Med. Chem. Anti-Cancer. Agents, 3, 95–117, 2003.<br />

93. Senger, D. R. et al., Vascular permeability factor (VPF, VEGF) in tumor biology, Metastasis Rev., 12,<br />

303–324, 1993.<br />

94. Dvorak, H. F., Brown, L. F., Detmar, M., and Dvorak, A. M., Vascular permeability factor/vascular<br />

endothelial growth factor, microvascular hyperpermeability and angiogenesis, Am. J. Pathol., 146,<br />

1029–1039, 1995.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


31<br />

CONTENTS<br />

Cell Penetrating Peptide (CPP)–<br />

Modified Liposomal<br />

Nanocarriers for Intracellular<br />

Drug and Gene Delivery<br />

Vladimir P. Torchilin<br />

31.1 Introduction ....................................................................................................................... 629<br />

31.2 Liposomes for Intracellular Drug Delivery ...................................................................... 630<br />

31.3 Transduction ...................................................................................................................... 631<br />

31.4 Intracellular Delivery of Pharmaceutical Nanocariers with CPPs ................................... 633<br />

31.5 Intracellular Delivery of Liposomes with CPPs ............................................................... 635<br />

References .................................................................................................................................... 637<br />

31.1 INTRODUCTION<br />

Although many important pharmacological targets for cancer therapy are located inside cells,<br />

membranes pose a serious obstacle to the uptake of charged hydrophilic molecules inside the<br />

cells. Many pharmacological proteins or peptides require to be intracellularly delivered to target<br />

and modulate cellular functions at the subcellular levels. In this connection, an important task is a<br />

proper identification of intracellular targets to be reached and affected. Cancer therapy provides<br />

a whole set of good examples to illustrate this importance. In a search of new anti-cancer drugs, a<br />

shift is gradually occurring from the semi-empirical approach based on evaluation of efficacy of<br />

candidate compounds against cultured cancer cells and animal tumor models to molecular<br />

mechanism-based drug discovery. 1 A huge body of information about cellular metabolic and<br />

signaling pathways essential for tumorogenesis and tumor cell development allows for the identification<br />

of proper targets for interference with the tumor growth. Quite a few molecular targets have<br />

already been identified. 1 The creation of a working draft of the human genome sequence 2,3 in<br />

combination with high-throughput methods of molecular biology promises continued rapid growth<br />

in identifying such targets. 4,5 However, not all drug targets that can be identified and validated <strong>by</strong><br />

molecular biology tools are considered suitable for drug development, often because of their<br />

intracellular location. 1 Myc/Max dimerization, Src homology-2 domain interaction, and Ras/Raf<br />

association have been identified as promising interference targets for the inhibition of tumor<br />

development. 6,7 Although peptide inhibitors, discovered through the use of phage-display 8 or<br />

combinatorial peptide 9 libraries had excellent in vitro activity, they were not considered for drug<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

629


630<br />

development because of poor pharmacokinetics and their inability to reach the molecular targets<br />

inside the cells. Sometimes, tumors result from malfunctions of tumor suppressors genes and the<br />

lack of activity of the proteins they encode. 5,10 In this case, the delivery into tumor cells of working<br />

copies of proteins obtained <strong>by</strong> recombinant methods would provide indispensable tools for validation<br />

of gene functions and potential development of protein or gene therapy-based methods of<br />

treatment. The use of these proteins for molecular target validation and eventual development of<br />

anti-cancer drugs is again hampered <strong>by</strong> low permeability of cell membranes.<br />

Intracellular transport of biologically active molecules with therapeutic properties is one of the<br />

key problems in drug delivery in general. However, the very nature of cell membranes prevents<br />

proteins, peptides, and nanoparticulate drug carriers from entering cells unless there is an active<br />

transport mechanism that is usually the case for very short peptides. 11 Various vector molecules<br />

promote the delivery of associated drugs and drug carriers inside the cells via receptor-mediated<br />

endocytosis. 12 This process involves the attachment of the vector molecule and an associated drug<br />

carrier to specific ligands on the target cell membranes, followed <strong>by</strong> the energy-dependent formation<br />

of endosomes. The problem, however, is that any molecule/particle entering cells via the<br />

endocytic pathway and becoming entrapped into endosome eventually ends in lysosome where<br />

active degradation processes take place under the action of numerous lysosomal enzymes. As a<br />

result, only a small fraction of unaffected substance appears in the cell cytoplasm. Even if an<br />

efficient cellular uptake via endocytosis is observed, the delivery of intact peptides and proteins<br />

is compromised <strong>by</strong> an insufficient endosomal escape and subsequent lysosomal degradation.<br />

Enhanced endosomal escape can be achieved through the use of, for example, lytic peptides, 13–<br />

15 pH-sensitive polymers, 16 or swellable dendritic polymers. 17 These agents have provided<br />

encouraging results in overcoming limitations of endocytosis-based cytoplasmic delivery, but<br />

there is still a need for further improvement or consideration of alternative delivery strategies.<br />

31.2 LIPOSOMES FOR INTRACELLULAR DRUG DELIVERY<br />

Nanotechnology for Cancer Therapy<br />

For almost two decades, liposomes have been considered promising carriers for biologically active<br />

substances. 18–20 They are biologically inert and completely biocompatible, and they cause practically<br />

no toxic or antigenic reactions. Drugs included into liposomes are protected from the<br />

destructive action of the external media, and liposomes are able to deliver their content inside<br />

cells and even inside different cell compartments. Water-soluble drugs can be captured <strong>by</strong> the inner<br />

water space of liposomes, whereas lipophilic compounds can be incorporated into the liposomal<br />

membrane. The definite drawback of liposomal preparations is their fast elimination from the blood<br />

and capture <strong>by</strong> the cells of the reticulo-endothelial system (RES), primarily, in the liver and spleen<br />

that is usually believed to be the result of rapid opsonization of the liposomes. 21 To increase<br />

liposome accumulation in the required areas, the use of targeted liposomes has been suggested. 22<br />

Liposomes with a specific affinity for an affected organ or tissue were believed to increase the<br />

efficacy of liposomal pharmaceutical agents. Immunoglobulins, primarily of the IgG class, and their<br />

fragments are the most promising and widely used targeting moieties for various drugs and drug<br />

carriers, including liposomes. Despite evident success in the development of antibody-to-liposome<br />

coupling techniques and improvements in the targeting efficacy, the majority of immunoliposomes<br />

still ended in the liver that was usually a consequence of insufficient time for the interaction<br />

between the target and targeted liposome.<br />

Different methods have been suggested to achieve long circulation of liposomes in vivo,<br />

including coating the liposome surface with inert, biocompatible polymers such as polyethylene<br />

glycol (PEG) that form a protective layer over the liposome surface and slows down the liposome<br />

recognition <strong>by</strong> opsonins and subsequent clearance. 23,24 Long-circulating liposomes are now widely<br />

used in biomedical in vitro and in vivo studies and even found their way into clinical practice. 25,26<br />

Explanations of the phenomenon involve the role of surface charge and hydrophilicity of PEG-coated<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 631<br />

liposomes, 27 the participation of PEG in the repulsive interactions between PEG-grafted membranes<br />

and other particles, 28 and more generally, the decreased rate of plasma protein adsorption on<br />

the hydrophilic surface of PEGylated liposomes. 29 Long-circulating immunoliposomes have also<br />

been prepared containing on their surface both antibody and PEG, there<strong>by</strong> possessing both abilities—<br />

i.e., to recognize and bind the target and to circulate long enough to provide high target<br />

accumulation. 30<br />

As previously discussed, most liposomes are internalized <strong>by</strong> cells via endocytosis and destined<br />

to lysosomes for degradation. 31 When one needs to achieve liposome-mediated drug delivery inside<br />

cells into a cytoplasmic compartment, pH-sensitive liposomes are frequently used (for one of many<br />

reviews, see V. P. Torchilin, F. Zhou, and L. Huang 1993). 32 As noticed in Puyal et al., 33 cellular<br />

drug delivery mediated <strong>by</strong> pH-sensitive liposomes is not a simple intracellular leakage from the<br />

lipid vesicle because the drug has to also cross the endosomal membrane. It is usually supposed that<br />

inside an endosome, the low pH and some other factors destabilize liposomal membrane that, in<br />

turn, interacts with the endosomal membrane, provoking its secondary destabilization and drug<br />

release into the cytoplasm. To prepare pH-sensitive immunoliposomes, the latter were additionally<br />

supplied with surface-immobilized antibodies. The advantages of antibody and pH-sensitive liposome<br />

combination are mutually additive: cytoplasmic delivery, targetability, and facilitated uptake<br />

(i.e., improved intracellular availability) via receptor-mediated endocytosis. Successful application<br />

of pH-sensitive immunoliposomes has been demonstrated in the delivery of a variety of molecules<br />

including fluorescent dyes, anti-tumor drugs, proteins, and DNA. 34<br />

31.3 TRANSDUCTION<br />

A promising approach that seems to be the solution of overcoming the cellular barrier has emerged<br />

over the last decade. In this approach, certain proteins or peptides can be tethered to the hydrophilic<br />

drug of interest, and together, the construct possesses the ability to translocate across the plasma<br />

membrane and intracellularly deliver the payload. This process of translocation is called protein<br />

transduction. Such proteins or peptides contain domains of less than 20 amino acids, termed as<br />

Protein Transduction Domains (PTDs) or CPPs, that are highly rich in basic residues. These<br />

peptides have been used for intracellular delivery of various cargoes with molecular weights<br />

several times greater than their own. 35 CPPs are a group of peptides, usually containing a cluster<br />

of basic residues that have been recognized as promising drug delivery vectors over the last decade.<br />

The CPPs are gaining increased attention as they possess the remarkable property of translocating<br />

across the hydrophobic cell membrane that forms a formidable barrier to the entry of hydrophilic<br />

and high molecular-weight drugs. As a result, different therapeutic moieties that are mostly hydrophilic<br />

in nature and/or are of high molecular weight can be tagged to CPPs and transported across<br />

the cell membrane to exert their pharmacological actions at the subcellular level. This process of<br />

traversing across the biological barrier is called protein transduction, and it is confined to a domain<br />

of less than 20 amino acids, termed as PTD or CPP. CPPs are rich in basic residues and are either<br />

derived from corresponding transducing proteins or synthesized. The original concept of<br />

protein transduction came out from the observation that the trans-activating transcriptional<br />

activator, TAT, protein encoded <strong>by</strong> HIV-1, was efficiently internalized <strong>by</strong> cells in vitro, resulting<br />

in trans-activation of the viral promoter. 36,37 Subsequently, several other proteins and peptides were<br />

found to display the translocation activity that encompasses penetratin, 38 VP22, 39 transportan, 40<br />

model amphipathic peptide MAP, 41 signal sequence-based peptides, 42 and synthetic arginineenriched<br />

sequences. 43<br />

Penetratin is a CPPs derived from the homeodomain of Antennapedia (Drosophila homeoprotein).<br />

Homeoproteins are transcription factors that comprise a stretch of 60 amino acids called the<br />

homeodomain that is involved in the DNA binding. The homeodomain was shown to traverse<br />

through the mammalian nerve cells and accumulate in their nuclei. 44 More specifically, the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


632<br />

Nanotechnology for Cancer Therapy<br />

translocation ability was narrowed down to a 16-mer peptide, termed as penetratin (Antp PTD, 43–<br />

58 residues, Arg-Gln-Ile-Lys-Ile-Trp-Phe-Gln-Asn-Arg-Arg-Met-Lys-Trp-Lys-Lys) present in the<br />

third helix of the homeodomain. 38 VP22 is a herpes virus type 1 protein associated with the<br />

transport between cells where it also ends up in the nuclei. 39 Transportan is a chimeric CPP<br />

built of galanin and mastoparan. Inside the cells, transportan is taken up into the nuclei and<br />

concentrates in subnuclear structures, probably the nucleoli. 40 MAP is a synthetic 18-mer<br />

peptide capable of transporting various cargoes across the cellular membrane. MAP has presented<br />

the highest uptake and cargo delivery efficiency among other CPPs. 45 Signal-sequence-based<br />

peptides comprise membrane translocating sequences (MTSs) that direct the pre-protein toward<br />

the accurate intracellular organelles. Such sequences, when coupled to nuclear localization signals,<br />

can be directed to accumulate in the nuclei of the cells. 46 To identify the specific regions of TAT<br />

peptide responsible for translocation, synthetic peptides rich in arginines were prepared such as<br />

arginine-substituted TAT (R9-Tat) and D-amino acid substituted TAT (D-TAT). 43 Such peptides<br />

were internalized with similar efficiencies as TAT peptide. In addition, various arginine-rich<br />

peptides such as RNA-binding peptides derived from proteins HIV-1 Rev, flock house virus<br />

coat, and DNA-binding peptides from c-Fos, c-Jun, and yeast GCN4 also displayed<br />

translocation properties.<br />

TAT peptide (TATp) remains the most frequently used CPP for drug delivery purposes. TAT is<br />

transcriptional activator protein encoded <strong>by</strong> human immunodeficiency virus type 1 (HIV-1). 47<br />

TAT-mediated transduction was first utilized in 1994 for the intracellular delivery of variety of<br />

cargos such as b-galactosidase, horseradish peroxidase, RNase A, and domain III of Pseudomonas<br />

exotoxin A in vitro. 48 In vivo transduction using TAT peptide (37–72)-conjugated to b-galactosidase<br />

resulted in protein delivery to different tissues such as the heart, liver, spleen, lung, and<br />

skeletal muscle. Next, attempts were made to narrow down to the specific domain responsible for<br />

transduction. For this, synthetic peptides with deletions in the a-helix domain and the basic cluster<br />

domain were prepared for investigating their translocation ability. 49 The whole basic cluster from<br />

48 to 60 residues was found accountable for membrane translocation because any deletions or<br />

substitutions of basic residues in TAT (48–60) reduced the cellular uptake property. Rothbard et<br />

al. 50 studied TAT (48–57) <strong>by</strong> deletion analysis. They found that deletion of Gly-48 did not affect the<br />

transduction efficiency, whereas deletions of Lys-50,51, Arg-55–57 and Gln-54 markedly reduced<br />

transduction efficiencies. Therefore, the minimal transduction domain was assigned to TAT (49–<br />

57) residues. The commonly studied transduction domain of TAT (TAT PTD) extends from<br />

residues 47–57: Tyr-Gly-Arg-Lys-Lys-Arg-Arg-Gln-Arg-Arg-Arg that contain six arginines<br />

(Arg) and two lysine residues. 51<br />

Two types of the endocytic uptake of the CPPs have been proposed: the classical clathrinmediated<br />

endocytosis and the lipid-raft-mediated caveolae endocytosis. Clathrin-mediated endocytosis<br />

involves the formation of clathrin-coated membrane pits that pinch off the membrane to<br />

form vesicles for subsequent processing. 52 This type of endocytosis was suggested in Console et al.<br />

where the TAT peptide showed the co-localization with the classical endocytic marker, transferrin.<br />

This was substantiated further in Lundberg, Wilkstrom, and Johansson where TAT PTD and Antp<br />

PTD showed uptake only at 378C. In addition, the internalization required the expression of<br />

negatively charged glycosaminoglycans on the cell surface for interaction with CPPs that was<br />

followed <strong>by</strong> endocytosis. Studies also suggested that the PTDs do not provoke a real translocation,<br />

but they are only responsible for cell surface adherence that subsequently results in their endocytosis<br />

and accumulation in endosomes. 53–57 In fact, direct electrostatic interaction between the<br />

positive residues of CPPs and the negative residues of the cell-surface proteoglycans or glycosaminoglycans<br />

(such as heparan sulfate, heparin) is required in internalization regardless of the<br />

mechanism of cellular uptake 54,58–60 although some studies suggested the lack of correlation<br />

between proteoglycans and transduction process. 61<br />

Another proposed mechanism for internalization is caveolae-mediated clathrin-independent<br />

uptake. Caveolae uptake involves the formation of flask-like uncoated invaginations (50–70 nm),<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 633<br />

principally composed of a subclass of detergent-resistant membrane domains enriched in cholesterol<br />

and sphingolipids called as lipid rafts. 62 This type of uptake was suggested in. 63,64 Unlike the<br />

rapid uptake of transferrin, a marker for clathrin-mediated endocytosis, the internalization of TATcargo<br />

was very slow, reaching the plateau after several hours with the co-localization of TATp with<br />

the markers of caveolar uptake. The cellular uptake was affected in the presence of drugs that either<br />

disrupt lipid rafts or alter caveolar trafficking.<br />

A different mechanism has been proposed for the transport of guanidinium-rich CPPs conjugated<br />

to small molecules (MW!3,000 Da). 65,66 The guanidinium groups of the CPPs form<br />

bidentate hydrogen bonds with the negative residues on the cell-surface; the resulting ion pairs<br />

then translocate across the cell membrane under the influence of the membrane potential. The ion<br />

pair dissociates on the inner side of the membrane, releasing the CPPs into the cytosol. The number<br />

of guanidinium groups is critical for translocation with around eight groups being the optimum<br />

number for the efficient translocation.<br />

A recent mechanism proposed for CPP-conjugated to large cargos (MWO30,000 Da) is<br />

nonclathrin, noncaveolar endocytosis, called macropinocytosis. Macropinocytosis is a nonspecific<br />

form of cellular uptake, brought about <strong>by</strong> large vesicles known as macropinosomes that are generated<br />

from the actin filaments. 67 Because a majority of fusion proteins remained entrapped in<br />

macropinosomes, the TAT transduction domain was conjugated to a fusogenic peptide, the<br />

N-terminus domain of the influenza virus hemagglutinin protein HA2, to trigger the release of<br />

the TAT-fusion protein from endosome, enhancing their nuclear transport. A very recent study also<br />

demonstrates the macropinocytosis mechanism for small PTD peptides (1,000–5,000 Da). 68<br />

Therefore, it looks like that more than one mechanism works for CPP-mediated intracellular<br />

delivery of small and large molecules. Individual CPPs or CPP conjugated to small molecules are<br />

internalized into cells via electrostatic interactions and hydrogen bonding, whereas CPP conjugated<br />

to large molecules occur via the energy-dependent macropinocytosis. However, in both cases, the<br />

direct contact between the CPPs and negative residues on cell-surface is a requisite for<br />

successful transduction.<br />

Because the majority of the studies believe in endocytosis as a major mechanism for transduction,<br />

an important moment is the escape of CPPs from the endosomes and their translocation to the<br />

nuclei. Studies suggest that endosomal acidification prior to the disruption is required for CPP<br />

escape. 69 Another study suggested that the TAT-fusion proteins enter cells via the endosomal<br />

pathway, circumvent lysosomal degradation, and then sequester in the periphery of the nuclei. 70<br />

Overall, the efficiency of nuclear translocation process is limited.<br />

31.4 INTRACELLULAR DELIVERY OF PHARMACEUTICAL<br />

NANOCARIERS WITH CPPs<br />

The applications of TAT-mediated delivery extend to nanoparticles delivery inside cells. The<br />

concept of TAT-mediated nanoparticle delivery was first realized in 1999 when dextran-coated<br />

superparamagnetic iron oxide nanoparticles (size ca. 40 nm) conjugated to TATp (48–57), showing<br />

significantly higher uptake in lymphocytes in vitro than control TAT-free particle. 71 The technique<br />

showed potential for magnetically labeling cells in order to allow for their magnetic separation or<br />

magnetic resonance (MR) imaging. In vivo, TATp effectively transduced iron nanoparticles into<br />

hematopoietic and neural progenitor cells for stem cell analysis 72 andintoT-cellsforMR<br />

imaging. 73 Contrary to nuclear translocation of TATp, particulate TATp-iron conjugates were<br />

observed in the cytoplasm and not in the nuclei. 74 Because superparamagnetic nanoparticles<br />

were shown to be useful MR contrast agents for imaging and for cell labeling and cell tracking,<br />

the conjugation of such nanoparticles with TATp provides better signal from the treated cells for<br />

MR imaging. 75 For in vivo MRI, cells need to be labeled with magnetic particles through internalizing<br />

receptors. The limitation, however, is that most of the cells lack efficient internalization<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


634<br />

Nanotechnology for Cancer Therapy<br />

receptors or pathways. This problem was overcome <strong>by</strong> attaching the CPP such as TAT peptide (48–<br />

57) to the dextran-coated superparamagnetic iron oxide particles (CLIO). The average size of the<br />

particles was 41 nm, and the conjugate carried average 6.7 TAT moieties per particle. The cellular<br />

uptake studies of TAT-CLIO were performed on mouse lymphocytes, human natural killer cells,<br />

and HeLa cells. In all the three cell lines tested, the uptake of TAT–CLIO nanoparticles was about<br />

100-fold higher than the nonmodified iron oxide particle. The labeling with TAT–CLIO did not<br />

induce toxicity and did not alter the differentiation or proliferation pattern of CD34C cells. The<br />

TAT–CLIO-labeled CD34C cells and control cells were subsequently intravenously injected into<br />

the immunodeficient mice. 75 Around 4% of the injected dose of the cells migrated to the bone<br />

marrow (per gram of the tissue), and the labeled and control cells showed similar biodistribution<br />

profile. Nonetheless, it was possible to visualize the labeled cells <strong>by</strong> MRI within mouse bone<br />

marrow at the single-cell level. Also, such magnetically labeled cells could be recovered from<br />

the bone marrow after in vivo homing using magnetic separation columns.<br />

Similarly, gold nanoparticles were also modified with TATp. 76 As with iron-conjugates, these<br />

particles did not reach the cell nuclei in experiments with NIH3T3 and HepG2 cells. The uptake of<br />

the particles was found to proceed <strong>by</strong> endocytosis. TATp (48–57) was also conjugated to FITCdoped<br />

silica nanoparticles (FSNPs) for bioimaging purposes. 77 TATp-modified nanoparticles have<br />

also been investigated for their capability to deliver the diagnostic and therapeutic agents across the<br />

blood–brain barrier. 78 TATp–FSNPs were prepared <strong>by</strong> microemulsion system and studied for<br />

labeling of human lung adenocarcinoma cells (A-549) in vitro. The cells were efficiently labeled<br />

with TAT–FSNPs, unlike FSNPs that showed no effective labeling. For in vivo bioimaging<br />

potential, TAT–FSNPs were intraarterially administered to the rats’ brains. TAT-conjugated<br />

FSNPs labeled the brain blood vessels, showing the potential for delivering agents to the brain<br />

without compromising the blood–brain barrier.<br />

A new application of CPP-mediated delivery is the labeling of cells with quantum dots using<br />

CPP-modified quantum dot-loaded polymeric micelles. 79 Quantum dots are gaining popularity over<br />

standard fluorophores for studying tumor pathophysiology because they are photostable and are<br />

very bright fluorophores. They can be tuned to a narrow emission spectrum, and they are relatively<br />

insensitive to the wavelength of the excitation light. Quantum dots have the ability to distinguish<br />

tumor vessels from both the perivascular cells and the matrix with concurrent imaging. Quantum<br />

dots were trapped within micelles prepared of PEG–phosphatidyl ethanolamine (PEG–PE) conjugates<br />

bearing TAT–PEG–PE linker. TAT-quantum dot conjugates could label mouse endothelial<br />

cells in vitro. For in vivo racking, bone marrow-derived progenitor cells were labeled with TATbearing<br />

quantum dot-containing micelle ex vivo and then injected in the mouse bearing tumor in a<br />

cranial window model. It was then possible to track the movement of labeled progenitor cell to<br />

tumor endothelium, introducing an attempt toward understanding fine details of tumor<br />

neovascularization.<br />

CPPs have also been used to enhance the delivery of genes via solid lipid nanoparticle (SLN). 80<br />

SLN gene vector was modified with dimeric TATp (TAT2) and compared with polyethylenimine<br />

(PEI) for gene expression in vitro and in vivo. The presence of TAT2 in SLN gene vector enhanced<br />

the gene transfection compared to PEI both in vitro and in vivo. In another study, TATp (47–57)<br />

conjugated to nanocage structures showed binding and transduction of the cells in vitro. 81 The shell<br />

cross-linked (SCK) nanoparticles were prepared <strong>by</strong> the micellization of amphiphilic block copolymers<br />

of poly(epsilon-caprolactone-b-acrylic acid) and conjugated to TATp that was<br />

independently built on a solid support, resulting in TAT-modified nanocage conjugate. Such<br />

conjugate was analyzed <strong>by</strong> confocal microscopy with CHO and HeLa cells. The conjugated nanoparticles<br />

showed binding and transduction inside the cells. The authors then characterized the SCK<br />

nanoparticles for the optimum number of TATp per particle required to enhance the transduction<br />

efficiency. 82 The authors prepared the conjugates with 52, 104, and 210 CPP peptides per particle<br />

that were then evaluated for the biocompatibility in vitro and in vivo. 83 In vitro studies showed the<br />

inflammatory responses to the conjugates, but in vivo evaluation of the conjugates in mice did not<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 635<br />

result in major incompatible responses. Therefore, TATp–SCK conjugate can be used as scaffolds<br />

for preparing antigen for immunization.<br />

TAT-mediated nanoparticulate delivery is also finding its way in vaccination fields to elicit<br />

better immune response. When TAT (1–72)-coated anionic nanoparticles were used to immunize<br />

mice, it generated antibodies and T helper type-1 immune response to TAT. 84 In another study,<br />

TAT microspheres were used for vaccinations. 85 Anionic microspheres of different compositions,<br />

size, and surface charge density were prepared, and all of them adsorbed biologically active TAT<br />

protein in a reversible mode. The microspheres were intracellularly delivered <strong>by</strong> TAT and were not<br />

toxic, both in vitro and in vivo.<br />

31.5 INTRACELLULAR DELIVERY OF LIPOSOMES WITH CPPs<br />

CPPs have also augmented the delivery of liposomal drug carriers. TAT peptide (47–57)-modified<br />

liposomes could be intracellularly delivered in different cells such as murine Lewis lung carcinoma<br />

(<strong>LLC</strong>) cells, human breast tumor BT20 cells, and rat cardiac myocyte H9C2 cells. 86 The liposomes<br />

were tagged with TAT peptide via the spacer p-nitrophenylcarbonyl–PEG–PE at the density of 500<br />

TAT peptide per single liposome vesicle. It was shown that the cells treated with liposomes where<br />

TAT peptide–cell interaction was hindered either <strong>by</strong> direct attachment of TAT peptide to the<br />

liposome surface or <strong>by</strong> the long PEG grafts on the liposome surface shielding the TAT moiety<br />

did not show TAT-liposome internalization; however, the preparations of TAT-liposomes that<br />

allowed for the direct contact of TAT peptide residues with cells displayed an enhanced uptake<br />

<strong>by</strong> the cells. This suggested that the translocation of TAT peptide (TATp)-liposomes into cells<br />

requires direct free interaction of TAT peptide with the cell surface. Further studies on the intracellular<br />

trafficking of rhodamine-labeled TATp-liposomes loaded with FITC-dextran revealed that<br />

TATp-liposomes remained intact inside the cell cytoplasm within 1 h of translocation; after 2 h,<br />

they migrated into the perinuclear zone, and eventually, the liposomes disintegrated there. 87<br />

The TATp-liposomes were also investigated for their gene delivery ability. For this, TATpliposomes<br />

prepared with the addition of a small quantity of a cationic lipid (DOTAP) were<br />

incubated with DNA. The liposomes formed firm noncovalent complexes with DNA. Such<br />

TATp-liposome-DNA complexes, when incubated with mouse fibroblast NIH 3T3 and cardiac<br />

myocytes H9C2, showed substantially higher transfection in vitro with lower cytotoxicity than<br />

the commonly used Lipofectin w . NIH/3T3, BT20, or H9C2 cells were transfected <strong>by</strong> incubation<br />

with TAT peptide–liposome/plasmid complexes in serum-free media for 4 h at 378C under 6%<br />

CO 2. The flow cytometry data demonstrated that the treatment of NIH/3T3 cells with TAT peptide–<br />

liposome/pEGFP-N1 complexes results in high fluorescence, i.e., high transfection outcome.<br />

Similar results were obtained with all cell lines tested. Confocal microscopy confirmed the transfection<br />

of both HCC150 and H9C2 cells with TAT peptide–liposome/DNA complexes<br />

(Figure 31.1). From 30 to 50% of both cell types in the field of view show a bright green fluorescence,<br />

whereas lower fluorescence was observed in virtually all cells. Under in vivo conditions,<br />

the intratumoral injection of TATp-liposome–DNA complexes into the <strong>LLC</strong> tumor in mice resulted<br />

in an efficient transfection of the tumor cells. Histologically, hematoxylin/eosin-stained tumor<br />

slices in both control and experimental animals showed a typical pattern of poorly differentiated<br />

carcinoma; however, under the fluorescence microscope, samples from control mice (nontreated<br />

mice or mice injected with TAT peptide–free liposome/plasmid complexes) showed only a background<br />

fluorescence, and slices from tumors injected with TAT peptide–liposome/plasmid<br />

complexes contained bright green fluorescence in tumor cells, indicating an efficient TAT<br />

peptide-mediated transfection. The study implicated the usefulness of TATp-liposomes for<br />

in vitro and localized in vivo gene therapy.<br />

Another study examined the kinetics of uptake of the TAT- and penetratin-modified liposomes.<br />

88 It was found that the translocation of liposomes <strong>by</strong> TAT peptide or penetratin was<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


636<br />

FIGURE 31.1 The enhancement of the liposome-mediated transfection <strong>by</strong> TAT peptide. HCC1500 cells<br />

(human breast carcinoma; a,b) or H9C2 cells (murine myoblasts; c) were incubated with the presence of<br />

liposome/pEGFP-N1 plasmid (encoding for the Green Fluorescence Protein, GFP) complexes (10 mg DNA per<br />

100,000 cells) for 4 h at 378C. The transfection efficiency (the appearance of the green fluorescence of the GFP<br />

inside cells) was detected after 72 h. (a), Cells were incubated with control plain (TAT-free) plasmid-bearing<br />

liposomes; (b,c), Cells were incubated with plasmid-bearing TAT-liposomes of the same composition. (a), (b)<br />

and (c)—fluorescent microscopy with FITC filter. Background transfection can only be seen with controls,<br />

whereas the introduction of TAT peptide into the preparation provides a dramatic enhancement of the GFP<br />

expression, i.e., increases the transfection efficiency.<br />

proportional to the number of peptide molecules attached to the liposomal surface. A peptide<br />

number of as few as five was already sufficient to enhance the intracellular delivery of liposomes.<br />

The kinetics of the uptake were peptide- and cell-type dependent. With TATp-liposomes, the<br />

intracellular accumulation was time-dependent, and with penetratin-liposomes, the accumulation<br />

within the cells was quick to reach the peak within 1 h, after that, it gradually declined.<br />

A study on the similar lines showed that Antp (43–58) and TAT (47–57) peptides coupled to<br />

small unilamellar liposomes were accumulated in higher proportions within tumor cells and<br />

dendritic cells than unmodified control liposomes. 89 The uptake was time- and concentrationdependent,<br />

and at least, 100 PTD molecules per small unilamellar liposomes were required for<br />

efficient translocation inside cells. The uptake of the modified liposomes was inhibited <strong>by</strong> the<br />

preincubation of liposomes with heparin, confirming the role of heparan sulfate proteoglycans in<br />

CPP-mediated uptake.<br />

In a different approach for improving the transfection and protecting DNA from degradation,<br />

thiocholesterol-based cationic lipids (TCL) were used in the formation of nanolipoparticles (NLPs).<br />

The NLPs were sequentially modified with TAT peptide that resulted in TAT–NLPs with a zwitterionic<br />

surface and higher transfection efficiency than for the cationic NLPs. 90<br />

TABLE 31.1<br />

Intracellular Delivery of Nanoparticles <strong>by</strong> CPPs<br />

Particle and size CPP Cell<br />

CLIO (MION) particles,<br />

40 nm<br />

Nanotechnology for Cancer Therapy<br />

TATp Mouse lymphocytes, human natural killer, HeLa, human hematopoietic<br />

CD34C, mouse neural progenitor C17.2, human lymphocytes CD4C,<br />

T-cells, B-cells, macrophages<br />

Gold particles, 20 nm TATp NIH3T3, HepG2, HeLa, human fibroblast HTERT-BJ1<br />

Quantum dot-loaded<br />

polymeric micelles,<br />

20 nm<br />

TATp Mouse endothelial cells, bone marrow-derived progenitor cells<br />

Sterically stabilized<br />

liposomes, 200 nm<br />

TATp Mouse <strong>LLC</strong>, human BT20, rat H9C2, <strong>LLC</strong> tumor in mice<br />

Sterically stabilized TATp or Human bladder carcinoma HTB-9, murine colon carcinoma C26, human<br />

liposomes, 65–75 nm penetratin epidermoid carcinoma A431, human breast cancer SK-BR-3, MCF7/WT,<br />

MCF7/ADR, murine bladder cancer MBT2, dendritic cells<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 637<br />

Although initial studies suggested an energy-independent character for the internalization of<br />

TATp-liposomes, 86 recent studies have revealed the endocytosis as the main mechanism for the<br />

intracellular uptake of TATp-liposomes. As was shown in M. M. Fretz, et al., 93 the conjugation of<br />

TAT peptide to lipoplexes enhanced the gene transfection in primary cell cultures <strong>by</strong> the endocytic<br />

uptake. Similarly, coupling of TAT peptide to the outer surface of liposomes resulted in an<br />

enhanced binding and endocytosis of the liposomes in ovarian carcinoma cells. 92 The binding<br />

was inhibited in the presence of heparin or dextran sulfate, suggesting that the proteoglycans<br />

expressed on the cell surface are also involved in cell binding in this case. In contrast, a new<br />

class of transducing peptides, haptides, after binding to the liposomal surface, augmented liposomes<br />

penetration through the cell membrane into the cell cytoplasm <strong>by</strong> a nonreceptor mediated<br />

process, 93 confirming that a variety of mechanisms could be involved in CPP-mediated intracellular<br />

delivery of nanoparticulates.<br />

CPPs clearly can serve as versatile delivery vectors for intracellular drug delivery. They can<br />

bring inside cells a wide range of cargos of different sizes—from small molecules to relatively large<br />

nanoparticles (see Table 31.1 for nanoparticles). Of the different CPPs used, TATP remains the<br />

most frequently used for this purpose in experimental cancer therapy. Intracellular cytoplasmic<br />

delivery of drugs and DNA <strong>by</strong> CPP-modified pharmaceutical nanocarriers may find applications for<br />

in vitro and ex vivo cell treatment as well as in different protocols of local drug application.<br />

REFERENCES<br />

1. Gibbs, J. B., Mechanism-based target identification and drug discovery in cancer research, Science,<br />

287, 1969–1973, 2000.<br />

2. Venter, J. C., Adams, M. D., Myers, E. W., Li, P. W., Mural, R. J., Sutton, G. G., Smith, H. O.,<br />

Yandell, M., Evans, C. A., and Holt, R. A., The sequence of the human genome, Science, 291,<br />

1304–1351, 2001.<br />

3. Lander, E. S., Linton, L. M., Birren, B., Nusbaum, C., Zo<strong>by</strong>, M. C., Baldwin, M.C, Devon, K., Dewar,<br />

K., Doyle, M., and Fitzhugh, W., Initial sequencing and analysis of the human genome, Nature, 409,<br />

860–921, 2001.<br />

4. Workman, P., New drug targets for genomic cancer therapy: Successes, limitations, opportunities and<br />

future challenges, Curr. Cancer Drug Targets, 1, 33–47, 2001.<br />

5. Balmain, A., Cancer genetics: From Boveri and Mendel to microarrays, Nat. Rev. Cancer, 1, 77–82,<br />

2001.<br />

6. Gibbs, J. B. and Oliff, A., Pharmaceutical research in molecular oncology, Cell, 79, 193–198, 1994.<br />

7. Sawyer, T. K., Src homology-2 domains: Structure, mechanisms, and drug discovery, Biopolymers,<br />

47, 243–261, 1998.<br />

8. Katz, B. A., Structural and mechanistic determinants of affinity and specificity of ligands discovered<br />

or engineered <strong>by</strong> phage display, Annu. Rev. Biophys. Biomol. Struct., 26, 27–45, 1997.<br />

9. Cortese, R., Ed, Combinatorial Libraries: Synthesis, Screening and Application Potential, Walter de<br />

Gruyter, Inc., New York, 1995.<br />

10. Hussain, S. P., Hofseth, L. J., and Harris, C. C., Tumor suppressor genes: At the crossroads of<br />

molecular carcinogenesis, molecular epidemiology and human risk assessment, Lung Cancer, 34<br />

(Suppl. 2), S7–S15, 2001.<br />

11. Egleton, R. D. and Davis, T. P., Bioavailability and transport of peptides and peptide drugs into the<br />

brain, Peptides, 18, 1431–1439, 1997.<br />

12. Park, J. W., Kirpotin, D. B., Hong, K., Shala<strong>by</strong>, R., Shao, Y., Nielsen, U. B., and Marks, J. D., Tumor<br />

targeting using anti-her2 immunoliposomes, J. Controlled Release, 74, 95–113, 2001.<br />

13. Kamata, H., Yagisawa, H., Takahashi, S., and Hirata, H., Amphiphilic peptides enhance the efficiency<br />

of liposome-mediated DNA transfection, Nucleic Acids Res., 22, 536–537, 1994.<br />

14. Midoux, P., Kichler, A., Boutin, V., Maurizot, J. C., and Monsigny, M., Membrane permeabilization<br />

and efficient gene transfer <strong>by</strong> a peptide containing several histidines, Bioconjugate Chem., 9, 260–267,<br />

1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


638<br />

Nanotechnology for Cancer Therapy<br />

15. Mastrobattista, E., Koning, G. A., Van Bloois, L., Filipe, A. C. S., Jiskoot, W., and Storm, G.,<br />

Functional characterization of an endosome-disruptive peptide and its application in cytosolic<br />

delivery of immunoliposome-entrapped proteins, J. Biol. Chem., 277, 27135–27143, 2002.<br />

16. Lackey, C. A., Press, O., Hoffman, A., and Styton, P., A biomimetic pH-responsive polymer directs<br />

endosomal release and intracellular delivery of an endocytosed antibody complex, Bioconjugate<br />

Chem., 13, 996–1001, 2002.<br />

17. Padilla De Jesus, O. L., Ihre, H. R., Gagne, L., Frechet, J. M., and Szoka, F. C. Jr., Polyester dendritic<br />

systems for drug delivery applications: In vitro and in vivo evaluation, Bioconjugate Chem., 13,<br />

453–461, 2002.<br />

18. Gregoriadis, G., Ed, Liposomes as Drug Carriers, Wiley, Chinchester, UK, 1988.<br />

19. Lasic, D. D., Liposomes. From Physics to Applications, Elsevier, Amsterdam, 1993.<br />

20. Lasic, D. D. and Papahadjopoulos, D., Eds., Medical Application of Liposomes, Elsevier, Amsterdam,<br />

1998.<br />

21. Senior, J. H., Fate and behavior of liposomes in vivo: A review of controlling factors, Crit. Rev. Ther.<br />

Drug Carrier Syst., 3, 123–193, 1987.<br />

22. Torchilin, V. P., Liposomes as targetable drug carriers, CRC Crit. Rev. Ther. Drug Carrier Syst., 1,<br />

65–115, 1985.<br />

23. Klibanov, A. L., Maruyama, K., Torchilin, V. P., and Huang, L., Amphipatic polyethyleneglycols<br />

effectively prolong the circulation time of liposomes, FEBS Lett., 268, 235–238, 1990.<br />

24. Torchilin, V. P. and Trubetskoy, V. S., Which polymers can make nanoparticulate drug carriers longcirculating?,<br />

Adv. Drug Deliv. Rev., 16, 141–155, 1995.<br />

25. Martin, F. and Lasic, D., Eds., Stealth w Liposomes, CRC Press, Boca Raton, 1995.<br />

26. Gabizon, A. A., Pegylated liposomal doxorubicin: Metamorphosis of an old drug into a new form of<br />

chemotherapy, Cancer Invest., 19, 424–436, 2001.<br />

27. Blume, G. and Cevc, G., Molecular mechanism of the lipid vesicle longevity in vivo, Biochim.<br />

Biophys. Acta, 1146, 157–168, 1993.<br />

28. Kenworthy, A. K., Hristova, K., Needham, D., and McIntosh, T. J., Range and magnitude of the steric<br />

pressure between bilayers containing phospholipids with covalently attached poly(ethylene glycol),<br />

Biophys. J., 68, 1921–1936, 1995.<br />

29. Woodle, M. C., Newman, M. S., and Cohen, J. A., Sterically stabilized liposomes: Physical and<br />

biological properties, J. Drug Target., 5, 397–403, 1994.<br />

30. Torchilin, V. P., Klibanov, A. L., Huang, L., O’Donnell, S., Nossiff, N. D., and Khaw, B. A., Targeted<br />

accumulation of polyethelene glycol-coated immunoliposomes in infarcted rabbit myocardium,<br />

FASEB J., 6, 2716–2719, 1992.<br />

31. Straubinger, R. M. N., pH-sensitive liposomes mediate cytoplasmic delivery of encapsulated macromolecules,<br />

FEBS Lett., 179, 148–154, 1985.<br />

32. Torchilin, V. P., Zhou, F., and Huang, L., pH-Sensitive liposomes (review), J. Liposome Res., 3,<br />

201–255, 1993.<br />

33. Puyal, C., Milhaud, P., Bienvenue, A., and Philippot, J. R., A new cationic liposome encapsulating<br />

genetic material. A potential delivery system for polynucleotides, Eur. J. Biochem., 228, 697–703,<br />

1995.<br />

34. Collins, D. and Huang, L., Cytotoxicity of diphtheria toxin A fragment to toxin-resistant murine cells<br />

delivered <strong>by</strong> pH-sensitive immunoliposomes, Cancer Res., 47, 735–739, 1987.<br />

35. Zorko, M. and Langel, U., Cell-penetrating peptides: Mechanism and kinetics of cargo delivery, Adv.<br />

Drug Deliv. Rev., 57, 529–545, 2005.<br />

36. Green, M. and Loewenstein, P. M., Autonomous functional domains of chemically synthesized<br />

human immunodeficiency virus tat trans-activator protein, Cell, 55, 1179–1188, 1988.<br />

37. Frankel, A. D. and Pabo, C. O., Cellular uptake of the tat protein from human immunodeficiency virus,<br />

Cell, 55, 1189–1193, 1988.<br />

38. Derossi, D., Joliot, A. H., Chassaing, G., Derossi, D., and Prochiantz, A., The third helix of the<br />

Antennapedia homeodomain translocates through biological membranes, J. Biol. Chem., 269,<br />

10444–10450, 1994.<br />

39. Elliot, G. and O’Harre, P., Intercellular trafficking and protein delivery <strong>by</strong> a herpesvirus structural<br />

protein, Cell, 88, 223–233, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 639<br />

40. Pooga, M., Hällbrink, M., Zorko, M., and Langel, Ü, Cell penetration <strong>by</strong> transportan, FASEB J., 12,<br />

67–77, 1998.<br />

41. Oehlke, J., Scheller, A., Wiesner, B., Krause, E., Beyermann, M., Klauschenz, E., Melzig, M., and<br />

Bienert, M., Cellular uptake of an alpha-helical amphipathic model peptide with the potential to<br />

deliver polar compounds into the cell interior non-endocytically, Biochim. Biophys. Acta, 1414,<br />

127–139, 1998.<br />

42. Lindgren, M., Hällbrink, M., Prochiantz, A., and Langel, U., Cell-penetrating peptides, Trends Pharmacol.<br />

Sci., 21, 99–103, 2000.<br />

43. Futaki, S., Suzuki, T., Ohashi, W., Yagami, T., Tanaka, S., Ueda, K., and Sugiura, Y., Arginine-rich<br />

peptides. An abundant source of membrane-permeable peptides having potential as carriers for<br />

intracellular protein delivery, J. Biol. Chem., 276, 5836–5840, 2001.<br />

44. Joliot, A., Pernelle, C., Deagostini-Bazin, H., and Prochiantz, A., Antennapedia homeobox peptide<br />

regulates neural morphogenesis, Proc. Natl. Acad. Sci. USA, 88, 1864–1868, 1991.<br />

45. Hällbrink, M., Florén, A., Elmquist, A., Pooga, M., Bartfai, T., and Langel, U., Cargo delivery kinetics<br />

of cell-penetrating peptides, Biochim. Biophys. Acta, 1515, 101–109, 2001.<br />

46. Lin, Y. Z., Yao, S. Y., Veach, R. A., Torgerson, T. R., and Hawiger, J., Inhibition of nuclear<br />

translocation of transcription factor NF-kappa B <strong>by</strong> a synthetic peptide containing a cell<br />

membrane-permeable motif and nuclear localization sequence, J. Biol. Chem, 270, 14255–14258,<br />

1995.<br />

47. Calnan, B. J., Tidor, B., Biancalana, S., Hudson, D., and Frankel, A. D., Arginine-mediated RNA<br />

recognition: The arginine fork, Science, 252, 1167–1171, 1991.<br />

48. Schwarze, S. R., Hruska, K. A., and Dowdy, S. F., Protein transduction: Unrestricted delivery into all<br />

cells?, Trends Cell Biol., 10, 290–295, 2000.<br />

49. Mitchell, D. J., Kim, D. T., Steinman, L., Fathman, C. G., and Rothbard, J. B., Polyarginine enters<br />

cells more efficiently than other polycationic homopolymers, J. Pept. Res., 56, 318–325, 2000.<br />

50. Rothbard, J. B., Kreider, E., Vandeusen, C. L., Wright, L., Wylie, B. L., and Wender, P. A., Argininerich<br />

molecular transporters for drug delivery: Role of backbone spacing in cellular uptake, J. Med.<br />

Chem., 45, 3612–3618, 2002.<br />

51. Sieczkarski, S. B. and Whittaker, G. R., Dissecting virus entry via endocytosis, J. Gen. Virol., 83,<br />

1535–1545, 2002.<br />

52. Richard, J. P., Melikov, K., Vives, E., Ramos, C., Verbeure, B., Gait, M. J., Chernomordik, L. V., and<br />

Lebleu, B., Cell-penetrating peptides. A reevaluation of the mechanism of cellular uptake, J. Biol.<br />

Chem., 278, 585–590, 2003.<br />

53. Console, S., Marty, C., Garcia-Echeverria, C., Schwendener, R., and Ballmer-Hofer, K., Antennapedia<br />

and HIV transactivator of transcription (TAT) protein transduction domains promote<br />

endocytosis of high molecular weight cargo upon binding to cell surface glycosaminoglycans,<br />

J. Biol. Chem., 278, 35109–35114, 2003.<br />

54. Lundberg, M., Wilkstrom, S., and Johansson, M., Cell surface adherence and endocytosis of protein<br />

transduction domains, Mol. Ther., 8, 143–150, 2003.<br />

55. Leifert, J. A., Harkins, S., and Whitton, J. L., Full-length proteins attached to the HIV tat protein<br />

transduction domain are neither transduced between cells, nor exhibit enhanced immunogenicity,<br />

Gene Ther., 9, 1422–14428, 2002.<br />

56. Vives, E., Richard, J. P., Rispal, C., and Lebleu, B., TAT peptide internalization: Seeking the<br />

mechanism of entry, Curr. Protein Pept. Sci., 4, 125–132, 2003.<br />

57. Tyagi, M., Rusnati, M., Presta, M., and Giacca, M., Internalization of HIV-1 tat requires cell surface<br />

heparan sulfate proteoglycans, J. Biol. Chem., 276, 3254–3261, 2001.<br />

58. Sandgren, S., Cheng, F., and Belting, M., Nuclear targeting of macromolecular polyanions <strong>by</strong> an HIV-<br />

Tat derived peptide. Role for cell-surface proteoglycans, J. Biol. Chem, 277, 38877–38883, 2002.<br />

59. Mai, J. C., Shen, H., Watkins, S. C., Cheng, T., and Robbins, P. D., Efficiency of protein transduction<br />

is cell type-dependent and is enhanced <strong>by</strong> dextran sulfate, J. Biol. Chem., 277, 30208–30218, 2002.<br />

60. Violini, S., Sharma, V., Prior, J. L., Dyszlewski, M., and Piwnica-Worms, D., Evidence for a plasma<br />

membrane-mediated permeability barrier to Tat basic domain in well-differentiated epithelial cells:<br />

Lack of correlation with heparan sulfate, Biochemistry, 41, 12652–12661, 2002.<br />

61. Anderson, R. G., The caveolae membrane system, Annu. Rev. Biochem., 67, 199–225, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


640<br />

Nanotechnology for Cancer Therapy<br />

62. Fittipaldi, A., Ferrari, A., Zoppe, M., Arcangeli, C., Pellegrini, V., Beltram, F., and Giacca, M., Cell<br />

membrane lipid rafts mediate caveolar endocytosis of HIV-1 Tat fusion proteins, J. Biol. Chem., 278,<br />

34141–34149, 2003.<br />

63. Ferrari, A., Pellegrini, V., Arcangeli, C., Fittipaldi, A., Giacca, M., and Beltram, F., Caveolaemediated<br />

internalization of extracellular HIV-1 tat fusion proteins visualized in real time, Mol.<br />

Ther., 8, 284–294, 2003.<br />

64. Rothbard, J. B., Jessop, T. C., and Wender, P. A., Role of membrane potential and hydrogen bonding<br />

in the mechanism of translocation of guanidinium-rich peptides into cells, J. Am. Chem. Soc., 126,<br />

9506–9507, 2004.<br />

65. Rothbard, J. B., Jessop, T. C., and Wender, P. A., Adaptive translocation: The role of hydrogen<br />

bonding and membrane potential in the uptake of guanidinium-rich transporters into cells, Adv.<br />

Drug Deliv. Rev., 57, 495–504, 2005.<br />

66. Swanson, J. A. and Watts, C., Macropinocytosis, Trends Cell Biol., 5, 424–428, 1995.<br />

67. Kaplan, I. M., Wadia, J. S., and Dowdy, S. F., Cationic TAT peptide transduction domain enters cells<br />

<strong>by</strong> macropinocytosis, J. Controlled Release, 102, 247–253, 2005.<br />

68. Potocky, T. B., Menon, A. K., and Gellman, S. H., Cytoplasmic and nuclear delivery of a TATderived<br />

peptide and a beta-peptide after endocytic uptake into HeLa cells, J. Biol. Chem., 278,<br />

50188–50194, 2003.<br />

69. Caron, N. J., Quenneville, S. P., and Tremblay, J. P., Endosome disruption enhances the functional<br />

nuclear delivery of Tat-fusion proteins, Biochem. Biophys. Res. Commun., 319, 12–20, 2004.<br />

70. Zhao, M., Kircher, M. F., Josephson, L., and Weissleder, R., Differential conjugation of tat peptide to<br />

superparamagnetic nanoparticles and its effect on cellular uptake, Bioconjugate Chem., 13, 840–844,<br />

2002.<br />

71. Lewin, M., Carlesso, N., and Tung, C. H., Tat peptide-derivatized magnetic nanoparticles allow<br />

in vivo tracking and recovery of progenitor cells, Nat. Biotechnol., 18, 410–414, 2000.<br />

72. Dodd, C. H., Hsu, H. C., Chu, W. J., Yang, P., Zhang, H. G., Zhang, J. D. Jr., Zinn, K., Forder, J.,<br />

Josephson, L., Weissleder, R., Mountz, J. M., and Mountz, J. D., Normal T-cell response and in vivo<br />

magnetic resonance imaging of T cells loaded with HIV transactivator-peptide-derived superparamagnetic<br />

nanoparticles, J. Immunol. Methods, 256, 89–105, 2001.<br />

73. Kaufman, C. L., Williams, M., Ryle, L. M., Smith, T. L., Tanner, M., and Ho, C., Superparamagnetic<br />

iron oxide particles transactivator protein-fluorescein isothiocyanate particle labeling for in vivo<br />

magnetic resonance imaging detection of cell migration: Uptake and durability, Transplantation,<br />

76, 1043–1046, 2003.<br />

74. Zhao, M., Kircher, M. F., Josephson, L., and Weissleder, R., Differential conjugation of tat peptide to<br />

superparamagnetic nanoparticles and its effect on cellular uptake, Bioconjugate Chem., 13, 840–844,<br />

2002.<br />

75. Tkachenko, A. G., Xie, H., Liu, Y. L., Coleman, D., Ryan, J., Glomm, W. R., Shipton, M. K., and<br />

Feldheim, D. L., Cellular trajectories of peptide-modified gold particle complexes: Comparison of<br />

nuclear localization signals and peptide transduction domains, Bioconjugate Chem., 15, 482–490,<br />

2004.<br />

76. Santra, S., Yang, H., Dutta, D., Stanley, J., Holloway, P., Tan, W., Moudgil, B., and Mericle, R., TAT<br />

conjugated, FITC doped silica nanoparticles for bioimaging applications, Chem. Commun. (Camb.),<br />

24, 2810–2811, 2004.<br />

77. Santra, S., Yang, H., Stanley, J. T., Holloway, P. H., Moudgil, B. M., Walter, G., and Mericle, R. A.,<br />

Rapid and effective labeling of brain tissue using TAT-conjugated CdS:Mn/ZnS quantum dots, Chem.<br />

Commun. (Camb.), 25, 3144–3146, 2005.<br />

78. Stroh, M., Zimmer, J. P., Zimmer, D. G., Levchenko, T. S., and Cohen, K. S., Quantum dots spectrally<br />

distinguish multiple species within the tumor milieu in vivo, Nat. Med., 11, 678–682, 2005.<br />

79. Rudolph, C., Schillinger, U., Ortiz, A., Tabatt, A., Plank, C., Müller, R. H., and Rosenecker, J.,<br />

Application of novel solid lipid nanoparticle (SLN)-gene vector formulations based on a dimeric<br />

HIV-1 TAT-peptide in vitro and in vivo, Pharm. Res., 21, 1662–1669, 2004.<br />

80. Liu, J., Zhang, Q., Remsen, E. E., and Wooley, K. L., Nanostructured materials designed for cell<br />

binding and transduction, Biomacromolecules, 2, 362–368, 2001.<br />

81. Becker, M. L., Remsen, E. E., Pan, D., and Wooley, K. L., Peptide-derivatized shell-cross-linked<br />

nanoparticles. 1. Synthesis and characterization, Bioconjugate Chem., 15, 699–709, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


CPP–Modified Liposomal Nanocarriers for Intracellular Drug and Gene Delivery 641<br />

82. Becker, M. L., Bailey, L. O., and Wooley, K. L., Peptide-derivatized shell-cross-linked nanoparticles.<br />

2. Biocompatibility evaluation, Bioconjugate Chem., 15, 710–717, 2004.<br />

83. Cui, Z., Patel, J., Tuzova, M., Ray, P., Phillips, R., Woodward, J. G., Nath, A., and Mumper, R. J.,<br />

Strong T cell type-1 immune responses to HIV-1 Tat (1-72) protein-coated nanoparticles, Vaccine, 22,<br />

2631–2640, 2004.<br />

84. Caputo, A., Brocca-Cofano, E., Castaldello, A., De Michele, R., Altavilla, G., Marchisio, M., Gavioli,<br />

R., Rolen, V., Chiarantini, L., Cerasi, A., Magnani, M., Cafaro, A., Sparnacci, K., Laus, M., Tondalli,<br />

L., and Ensoli, B., Novel biocompatible anionic polymeric microspheres for the delivery of the HIV-1<br />

Tat protein for vaccine application, Vaccine, 22, 2910–2924, 2004.<br />

85. Torchilin, V. P., Rammohan, R., Weissig, V., and Levchenko, T. S., TAT peptide on the surface of<br />

liposomes affords their efficient intracellular delivery even at low temperature and in the presence of<br />

metabolic inhibitors, Proc. Natl. Acad. Sci. USA, 98, 8786–8796, 2001.<br />

86. Torchilin, V. P., Levchenko, T. S., Rammohan, R., and Volodina, B., Cell transfection in vitro and<br />

in vivo with nontoxic TAT peptide-liposome-DNA complexes, Proc. Natl. Acad. Sci. USA, 100,<br />

1972–1977, 2003.<br />

87. Tseng, Y. L., Liu, J. J., and Hong, R. L., Translocation of liposomes into cancer cells <strong>by</strong> cellpenetrating<br />

peptides penetratin and tat: A kinetic and efficacy study, Mol. Pharmacol., 62,<br />

864–872, 2002.<br />

88. Marty, C., Meylan, C., Schott, H., Ballmer-Hofer, K., and Schwendener, R. A., Enhanced heparan<br />

sulfate proteoglycan-mediated uptake of cell-penetrating peptide-modified liposomes, Cell Mol. Life<br />

Sci., 61, 1785–1794, 2004.<br />

89. Huang, Z., Li, W., MacKay, J. A., and Szoka, F. C. Jr., Thiocholesterol-based lipids for ordered<br />

assembly of bioresponsive gene carriers, Mol. Ther., 11, 409–417, 2005.<br />

90. Hyndman, L., Lemoine, J. L., Huang, L., Porteous, D. J., and Boyd, A. C., HIV-1 Tat protein<br />

transduction domain peptide facilitates gene transfer in combination with cationic liposomes,<br />

J. Controlled Release, 99, 435–444, 2004.<br />

91. Fretz, M. M., Koning, G. A., Mastrobattista, E., Jiskoot, W., and Storm, G., OVCAR-3 cells internalize<br />

TAT-peptide modified liposomes <strong>by</strong> endocytosis, Biochim. Biophys. Acta, 1665, 48–56, 2004.<br />

92. Gorodetsky, R., Levdansky, L., Vexler, A., Shimeliovich, I., Kassis, I., Ben-Moshe, M., Magdassi, S.,<br />

and Marx, G., Liposome transduction into cells enhanced <strong>by</strong> haptotactic peptides (Haptides) homologous<br />

to fibrinogen C-termini, J. Controlled Release, 95, 477–488, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


32<br />

CONTENTS<br />

RGD-Modified Liposomes<br />

for Tumor Targeting<br />

P. K. Dubey, S. Mahor, and S. P. Vyas<br />

32.1 Introduction ....................................................................................................................... 643<br />

32.2 Tumor Vasculature Targeting ........................................................................................... 644<br />

32.2.1 Rationale for Tumor Vasculature Targeting....................................................... 644<br />

32.2.2 Molecular Targets in Angiogenic Tumor Vasculature....................................... 646<br />

32.3 Role of RGD in Integrin- and Fibronectin-Mediated Bioevents...................................... 647<br />

32.3.1 Integrins ............................................................................................................... 647<br />

32.3.2 Fibronectin........................................................................................................... 650<br />

32.3.3 RGD Mediated Metastasis Inhibition ................................................................. 651<br />

32.4 RGD-Modified Liposomes for Cancer Therapeutics........................................................ 653<br />

32.5 RGD-Modified Liposomes in Cancer Gene Therapy ....................................................... 656<br />

References .................................................................................................................................... 657<br />

32.1 INTRODUCTION<br />

Various drug delivery systems have been developed or are under development in order to minimize<br />

drug degradation or loss, to prevent harmful side effects, and to increase drug bioavailability and the<br />

fraction of the drug accumulated in the required zone. Among drug carriers, one can name microparticles,<br />

nanoparticles, synthetic polymers, cell ghosts, lipoproteins, micelles, niosomes, and<br />

liposomes. Each of these carriers offers its own advantages and has its own shortcomings; therefore,<br />

the choice of a certain carrier for each given case can be made only taking into account the<br />

whole bunch of relevent considerations. These carriers can be made slowly biodegradable,<br />

stimuli-sensitive, e.g., pH- or temperature sensitive, and even targeted, e.g., <strong>by</strong> conjugating them<br />

with various ligands.<br />

Liposomes-encapsulated anti-cancer drugs reveal their potential for increased therapeutic efficacy<br />

and decreased nonspecific toxicities as a result of their ability to enhance the delivery of<br />

chemotherapeutic agents selectively or preferentially to tumors (Papahadjopoulos and Gabizon<br />

1995; Martin 1998). However, these liposomes localize in the tumor extracellular compartment<br />

and are rarely taken up <strong>by</strong> the cancer cells located deep within the tumor (Papahadjopoulos et al.<br />

1991a, 1991b; Gabizon and Papahadjopoulos 1992) so that these cancer cells are readily reached<br />

with high concentration of drug and are given an opportunity to opt for drug-resistance. To further<br />

enhance cytotoxic efficacy, selective delivery of drugs to target cells can be achieved through<br />

liposomes appended with antibodies (Maruyama et al. 1990; Lopes de Menezes, Pilarski, and<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

643


644<br />

Allen 1998), serum proteins (Brown and Silvius 1990; Lundberg, Hong, and Papahadjoppoulos<br />

1993), or antibody fragments (Park et al. 1995; Kirpotin et al. 1997) that categorically recognize<br />

specific determinants (receptors) on target cells. However, attachment of antibodies and larger<br />

proteins to the exterior of liposomes has not been without problems. Chemistries producing phospholipid<br />

headgroup-protein coupling are often complex, jeopardizing the subsequent use of such<br />

engineered liposomes, especially in living systems. These chemistries can also change the protein’s<br />

confirmation and do not guarantee the final orientation with respect to either the potential ligand or<br />

the phospholipid bilayer, there<strong>by</strong> affecting the effectiveness of the targeting system.<br />

An alternative to the coupling of larger proteins is an anchoring of a defined small peptide<br />

domain on to the liposome surface (Gyongyossy-Issa, Muller, and Devine 1998). Ligand targeting<br />

using small peptides has advantages over the use of large protein molecules such as antibodies.<br />

These include ease of preparation, potentially lower antigenicity, and increased stability (Forssen<br />

and Willis 1998). RGD peptides have reported as effective in delivering cytotoxic molecules to the<br />

tumor vasculature (Arap, Pasqualini, and Ruoslahti 1998; Eller<strong>by</strong> et al. 1999; Gerlag et al. 2001;<br />

Schiffelers et al. 2003).<br />

32.2 TUMOR VASCULATURE TARGETING<br />

There are many potential barriers to the effective delivery of chemotherapeutic agents, having<br />

narrowest therapeutic indices in all medicines to the disseminated malignant tumors. The nonselective<br />

toxic effects on normal tissues restrict the dose of the anti-cancer agents. Through selective<br />

delivery of chemotherapeutic agents into the malignant tumors, the concentration of drugs in the<br />

tumors could be increased. This results in a decrease in the amount and types of nonspecific<br />

toxicites and an increased amount of drug that can be effectively delivered to the tumor. Monoclonal<br />

antibodies have been used to provide existing targeting opportunities. However, this<br />

approach has met with limited success; only a few antigens are known, and their expression on<br />

the cells within an individual tumor is not necessarily uniform. Moreover, antibodies against tumor<br />

antigens poorly penetrate into solid tumors (Dvorak et al. 1991). Furthermore, antibody targeted<br />

drug delivery may be impaired <strong>by</strong> clonal selection for cells that have lost the tumor antigen because<br />

tumor cells are genetically unstable and adopt for mutations advantageous for growth. The targeting<br />

of drug delivery systems to the tumor vasculature overcomes some of the problems associated with<br />

conventional tumor targeting.<br />

Angiogenesis is an important process in tumor growth as well as in metastasis. Growth and<br />

survival of tumor cells depend on oxygen and nutrients supplied <strong>by</strong> the blood. As a consequence,<br />

the size of a tumor is restricted to a few cubic millimeters without the recruitment of new blood<br />

vessels. The architecture of a solid tumor is such that numerous layers of tumor cells are fed <strong>by</strong> one<br />

blood vessel. This poses a significant barrier for selective delivery of cytotoxic drugs into malignant<br />

tumors that need to extravasate from blood into the tumor tissue to reach the target cell; the larger<br />

the chosen carrier, the less accesible the tumor tissue will be. In contrast, endothelial cells lining the<br />

tumor vasculature are easily accessible for these macromolecular preparations. The tumor vasculature<br />

targeting is a valuable tool for selective delivery of drugs is which either do not reach the<br />

target cells insufficiently or are too toxic to non-target cells when administered systemically.<br />

Figure 32.1 shows the principle of targeted chemotherapy to tumor blood vessels.<br />

32.2.1 RATIONALE FOR TUMOR VASCULATURE TARGETING<br />

Nanotechnology for Cancer Therapy<br />

Vascular targeting offers several benefits over conventional targeting to tumor cells (Table 32.1).<br />

The endothelium of the tumor vasculature is readily accessible to a circulating probe, whereas a<br />

conventional tumor targeting probe has to overcome the high interstitial pressure within the solid<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 645<br />

Receptors overexpressed<br />

on tumor<br />

vasculature<br />

Carrier constructs<br />

bearing antibody or<br />

ligands<br />

Blood vessel<br />

Tumor<br />

tissue<br />

FIGURE 32.1 Principle of targeted chemotherapy to tumor blood vessels.<br />

No<br />

penetration<br />

Normal tissue<br />

Peptide<br />

Liposome<br />

tumor <strong>by</strong> penetrating among closely packed tumor cells and dense tumor stroma (Dvorak et al.<br />

1991).<br />

The survival and growth of the tumor cells largely depend on blood supply to them. An anticancer<br />

therapy directed against the tumor vasculature does not have to eliminate every endothelial<br />

×<br />

Polyethylene glycol chain<br />

TABLE 32.1<br />

Comparison Between Conventional Tumor Targeting and Vascular Tumor Targeting<br />

Feature Conventional Tumor Targeting Vascular Tumor Targeting<br />

Pharmacokinetic system Multiple compartment Single compartment<br />

Primary target cell type Malignant tumor cells Non-malignant activated tumor<br />

vascular cells<br />

Target cell accessibility Poor Good<br />

Receptor Tumor antigen Angiogenic markers<br />

Receptor expression Uneven and heterogeneous Possibly uniform and homogeneous<br />

Homing moiety Monoclonal antibodies Monoclonal antibodies, peptides<br />

Therapeutic moiety Cytotoxic drug Cytotoxic drug, angiogenesis<br />

inhibitors<br />

Target cell genome Genetically unstable cells Genetically stable, diploid cells<br />

Acquired cytotoxic drug resistance Yes Not yet reported<br />

Posttargeting amplification loop Absent Present<br />

Complete elimination of target cell Required Not required<br />

Potential clinical applicability Limited <strong>by</strong> antigen heterogeneity Broad, as angiogenic markers may be<br />

shared<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


646<br />

cell; partial denuding of the endothelium is likely to lead to the formation of an occlusive thrombus<br />

that stops flow to the part of the tumor served <strong>by</strong> the vessel. In addition, tumor vasculature targeting<br />

has an intrinsic amplification mechanism. It has been estimated that elimination of a single endothelial<br />

cell can inhibit the growth of approximately 100 tumor cells (Denekamp 1993). Moreover,<br />

tumor endothelial cells are diploid and non-malignant, and they are unlikely to lose cell surface<br />

target receptors or acquire drug resistance through mutation and clonal evolution.<br />

Oncologists have long recognized that tumors commonly develop resistance to chemotherapy,<br />

whereas normal tissues do not. Therefore, toxicity to normal tissues such as chemotherapy-induced<br />

myelosuppression continues to occur even after tumor cells have become drug resistant and<br />

progress in the face of continued treatment. Endothelial cells, being non-malignant cells, are<br />

expected to behave in a manner analogous to bone marrow cells. Long-term antiangiogenic<br />

therapy has not been shown to produce drug resistance in experimental animals (Boehm et al.<br />

1997) or in clinical trials (Folkman 1997).<br />

32.2.2 MOLECULAR TARGETS IN ANGIOGENIC TUMOR VASCULATURE<br />

For successful targeting to angiogenesis-associated blood vessels, angiogenic endothelial cells need<br />

to be discriminated from the normal quiescent endothelium. In the past decade, using cellular and<br />

molecular biological approaches, many receptors, e.g., a vb 3 (vitronectin receptor), VEGFR, MMP-<br />

2/-9, endoglin (CD105), and aminopeptidase N (CD13), have been identified to be differentially<br />

upregulated on tumor endothelial cells. These receptors can be explored as target determinants for<br />

drug targeting purposes (Table 32.2).<br />

Several approaches exploiting different target epitopes/determinants were investigated for their<br />

potential to interfere with the endothelial neovascularization. VEGFR-I and VEGFR-2 are overexpressed<br />

on tumor vasculature, while being present at a low density in the surrounding normal<br />

tissues. VEGF-dephtheria toxin conjugate treatment of tumor bearing mice resulted in selective<br />

vascular damage in the tumor tissue and inhibited tumor growth (Olson et al. 1997). Histological<br />

studies revealed that conjugate treatment spared the blood vessels of normal tissues such as liver,<br />

lung, and kidney from being damaged.<br />

Endoglin (CDI05) is a transmembrane glycoprotein, over-expressed in the vasculature of<br />

tumors and other tissues undergoing vascular remodelling (Miller et al. 1999). Differential<br />

TABLE 32.2<br />

Potential Target Epitopes on Tumor Vascular Endothelium<br />

Target Epitope Reference<br />

30.5-kDa antigen Hagemeier et al. (1986)<br />

CD34 Schlingemann et al. (1990)<br />

VEGF/VEGF receptor complex Ramakrishnan et al. (1996)<br />

VEGF receptor Dvorak et al. (1991)<br />

Endosialin Rettig et al. (1992)<br />

E-selectin Nguyen et al. (1993)<br />

aV integrins Brooks et al. (1994)<br />

Endoglin Burrows et al. (1995)<br />

Tie-2 Sato et al. (1995)<br />

TNF a receptor Eggermont et al. (1996)<br />

CD44 Griffioen et al. (1997)<br />

Angiostatin receptor Moser et al. (1999)<br />

Endostatin receptor Chang et al. (1999)<br />

MMP-2/MMP-9 Koivunen et al. (1999)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


RGD-Modified Liposomes for Tumor Targeting 647<br />

up-regulation of endoglin presents an interesting opportunity to the selective delivery of cytotoxic<br />

molecules to the target endothelial cells. Several monoclonal antibodies specific for human endoglin<br />

have been produced (Seon et al. 1997). Radioiodinated monoclonal antibodies (10 mi) given to<br />

tumor-bearing animals significantly inhibited the tumor growth. This indicates the clinical potential<br />

of cytotoxic agent targeting using endoglin (Tabata et al. 1999).<br />

Other receptors over-expressed in tumor vessels include FGF receptors and Tie-2 receptors.<br />

Davol and colleagues prepared an endothelial cell-specific cytotoxic conjugate (Davol et al. 1995)<br />

<strong>by</strong> chemically linking the plant-derived ribosomal inhibitory protein saporin to FGF. The FGFsaporin<br />

conjugate inhibited proliferation of endothelial cells effectively in vitro. Radioimmunotherapy<br />

using 213 Bi or 131 I targeted to lung vasculature with a monoclonal antibody against<br />

thrombomodulin has been reported (Kennel and Mirzadeh 1998). Because thrombomodulin is<br />

selectively expressed in the pulmonary blood vessels, the antibody homed to lungs resulted in<br />

destruction of small tumor colonies. Cationic liposomes were used to target angiogenic vasculature<br />

in mouse pancreatic islet cell tumors. Confocal microscopy demonstrated a 15- to 33-fold preferentially<br />

higher uptake of the liposomes <strong>by</strong> angiogenic than normal endothelial cells (Thurston et al.<br />

1998).<br />

Erkki Ruoslahti and colleagues (1996) developed a novel targeting strategy <strong>by</strong> using polypeptidal<br />

system capable of delivering cytotoxic drugs selectively to integrins. An in vivo selection<br />

of phage display libraries identified peptides that specifically home to tumor blood vessels.<br />

Ruoslahti’s research group identified two major classes of peptides, one containing the RGD<br />

motif and the other containing an NGR motif. These polypeptides were then chemically linked<br />

to the anti-cancer drug doxorubicin. Treatment of breast carcinoma-bearing mice with the conjugated<br />

doxorubicin caused selective vascular damage in the tumors with consequential strong antitumor<br />

effect at a 10–40 times lower concentration as compared of free doxorubicin, whereas liver<br />

and heart toxicity was significantly reduced compared to that observed with free doxorubicin (Arap,<br />

Pasqualini, and Ruoslahti 1998). Their results illustrate the potential of targeting therapeutic agents<br />

to integrins expressed on the vasculature of tumors as an effective means of cancer treatment. Other<br />

approaches have also been used, e.g., toxic drugs (Arap, Pasqualini, and Ruoslahti 1998; Arora<br />

et al. 1999), apoptosis-inducing agents (Eller<strong>by</strong> et al. 1999), cytokines (Curnis et al. 2000), and<br />

therapeutic genes (Hood et al. 2002) were delivered to the tumor endothelial cells, leading to<br />

reduced tumor growth. Furthermore, angiogenesis and tumor growth were inhibited upon occlusion<br />

of blood vessels <strong>by</strong> targeting coagulation factors to tumor vasculature (Huang et al. 1997; Ran et al.<br />

1998). Recently, immune effector cells have successfully been targeted to angiogenic blood vessels,<br />

with resultant lysis of tumor endothelial cells and suppression of tumor growth (Niederman et al.<br />

2002).<br />

32.3 ROLE OF RGD IN INTEGRIN- AND FIBRONECTIN-MEDIATED BIOEVENTS<br />

Among the many molecules that are selectively expressed in angiogenic vasculature, the integrins<br />

show particular promise in tumor targeting. Saiki and co-workers reported that tumor angiogenesis<br />

can be inhibited <strong>by</strong> blocking the interaction between integrins and the RGD motif-containing extra<br />

cellular matrix proteins (Saiki et al. 1990). One of the most studied target epitopes on angiogenic<br />

neovasculature is the avb3 integrin. Studies show that avb3 and avb5 integrins are upregulated in<br />

angiogenic endothelial cells (Brooks, Clark, and Cheresh, 1994a) and the inhibition of these<br />

integrins <strong>by</strong> antibodies, cyclic RGD peptides (Brooks et al. 1994b) and RGD peptidomimetics<br />

(Carron et al. 1998) can block neovascularization.<br />

32.3.1 INTEGRINS<br />

Members of the integrin family of adhesion molecules are non-covalently-associated a/b heterodimers<br />

that mediate cell–cell, cell–extracellular matrix, and cell–pathogen interactions <strong>by</strong> binding<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


648<br />

to distinct, but often overlapping, combinations of ligands. Eighteen different integrin a-subunits<br />

and eight different b-subunits have been identified in vertebrates that form at least 24 a/b heterodimers,<br />

perhaps making integrins the most structurally and functionally diverse family of celladhesion<br />

molecules (Hynes 1992; Springer 1994) (Figure 32.2). Half of integrin a-subunits<br />

contains inserted (I) domains that are the principal ligand-binding domains (Shimaoka 2002).<br />

The complexity and structural and functional diversity of integrins allow this family of adhesion<br />

molecules to play a pivotal role in broad contexts of biology, including inflammation, innate and<br />

antigen specific immunity, haemostasis, wound healing, tissue morphogenesis, and regulation of<br />

cellular growth and differentiation. Conversely, disregulation of integrins is involved in the pathogenesis<br />

of many diseased states from autoimmunity to thrombotic vascular diseases to cancer<br />

metastasis (Curley, Blum, and Humphries 1999). Therefore, extensive efforts have been directed<br />

toward the discovery and development of integrins antagonists for clinical applications.<br />

Significant advances have been made in targeting aIIbb3 integrins on platelets for inhibiting<br />

thrombosis (Scarborough and Gretler 2000), avb3 and avb5 for blocking tumor metastasis, angiogenesis<br />

and bone resorption (Varner and Cheresh 1996) and b2 integrins and a4 integrins on<br />

leukocytes for treating autoimmune diseases and other inflammatory disorders (Giblin and Kelly<br />

2001; Yusuf-Makagiansar et al. 2002). Small-molecules, i.e., integrin inhibitors not only interfere<br />

with ligand binding but also stabilize particular integrin conformations that have provided insights<br />

into integrin structural rearrangements.<br />

The general structural features of all integrins appear to be similar. Both a- and b-subunits are<br />

transmembrane glycoproteins with large globular amino-terminal extracellular domains that,<br />

together, make up an ellipsoidal head (Figure 32.3). Each subunit provides a relatively thin leg<br />

that traverses across the plasma membrane and ends into a relatively small cytoplasmic tail of less<br />

than 60 amino acids. The only known integrin that does not fit this general description is b4 integrin<br />

that has a cytoplasmic domain of close to 1000 amino acids (Suzuki and Naitoh 1990; Tamura et al.<br />

1990).<br />

αL<br />

αM<br />

αX<br />

αD<br />

αE β7<br />

β4<br />

α1<br />

α1<br />

0<br />

α1<br />

1<br />

α2<br />

α3<br />

α4<br />

α5<br />

α6<br />

α7<br />

α8<br />

α9<br />

FIGURE 32.2 Integrin heterodimer composition. Integrin a- and b-subunits form 24 heterodimers that recognize<br />

distinct, but overlapping, ligands. (Adapted from Hynes, R. O., Cell, 69, 11, 1992.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

β1<br />

Nanotechnology for Cancer Therapy<br />

αv<br />

β3<br />

β5<br />

β6<br />

β8<br />

αII<br />

b


RGD-Modified Liposomes for Tumor Targeting 649<br />

26−28 nm<br />

4 nm<br />

8 nm<br />

12−14 nm<br />

S - S<br />

2 nm<br />

a b<br />

6 nm<br />

-S-S-<br />

The b1 integrins that can bind fibronectin include a3b1, a4b1, a5b1, and avb1. Many integrins,<br />

including the a3b1, a5b1, avb1, avb3, avb5, avb6, and the aIIbb3, recognize the RGD site located in<br />

adhesive proteins (Table 32.3). The fibronectin-specific integrin that consists of an a5 subunit and a<br />

b 1 subunit is the major fibronectin receptor expressed on most of the cells. This integrin mediates<br />

such cellular responses to fibronectin as adhesion, migration, assembly of a cytoskeleton, and<br />

assembly of the fibronectin extracellular matrix (The a 5b 1 integrin interacts with the central cell<br />

adhesive region of fibronectin and requires both the RGD and synergy sites for maximal binding<br />

(Obara and Yoshizato 1995). The major platelet integrin a IIhb 3 also recognizes a similar synergy<br />

site in fibronectin for mediating platelet interactions with fibronectin (Bowditch et al. 1994).<br />

Ca 2+<br />

Ca 2+<br />

Ca 2+<br />

Ca 2+<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

-S-S-<br />

FIGURE 32.3 The structural model of a5b1 integrin adhesion molecule. A structural model based on<br />

predicted primary structure and electron microscopy visualization of purified integrin is shown. The a5subunit<br />

consists of part of the globular extracellular domain and one leg that contains predicted 12 b strands<br />

(gray box). The b1- subunit makes up the remainder of the globular extracellular domain. The b1- subunit also<br />

contains five cysteine-rich repeats. Both the a5- and b1- subunits have transmembrane domains and small<br />

cytoplasmic domains. (Adapted from Varner, J. A. and Cheresh, D. A., Important Adv. Oncol., 69, 1996.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


650<br />

TABLE 32.3<br />

Integrin Receptors and Their Binding Sites<br />

Receptor Ligand<br />

32.3.2 FIBRONECTIN<br />

Receptor-Mediated<br />

Action<br />

Nanotechnology for Cancer Therapy<br />

Amino Acid Sequence<br />

Recognized<br />

a2b1 Collagen Adhesion DGEA<br />

a5b1 Fibronectin Adhesion RGD<br />

a 6b 1 Laminin Adhesion YIGSR<br />

aIIbb3 Fibrinogen Aggregation KQAGDV or RGD<br />

Fibronectin Aggregation RGD<br />

von Willebrand factor Aggregation RGD<br />

Vitronectin Aggregation RGD<br />

a6b3 Vitronectin Adhesion RGD<br />

Fibrinogen Adhesion RGD<br />

Fibronectinvon Adhesion RGD<br />

von Willebrand factor Adhesion RGD<br />

Fibronectin is a large adhesive glycoprotein found in extracellular matrices and body fluids<br />

(Mosher 1989; Carsons 1990; Hynes 1990). The primary structure of fibronectin is comprised of<br />

three different types of homologous repeating units or modules (Figure 32.4). There are several<br />

alternatively spliced forms of fibronectin that result from the deletion or insertion of complete type<br />

III modules. In addition, one particular region designated as IIICS can be partially inserted or<br />

deleted in some isoforms <strong>by</strong> alternative splicing. The homologous modules comprising of fibronectin<br />

are arranged in protease-resistant domains that are separated <strong>by</strong> more flexible, proteasesusceptible<br />

regions. When cleaved from intact fibronectin <strong>by</strong> partial proteolysis or expressed in<br />

bacterial or mammalian cells and individually purified, these domains often retain the specific<br />

binding functions of intact fibronectin such as those for he-parin, fibrin, denatured collagen<br />

(gelatin), and cell surface receptors.<br />

Fibronectin contains at least two distinct regions that can independently interact with distinct<br />

cell surface receptors. The first fibronectin cell-adhesive site to be identified was isolated in the<br />

form of protease-resistant fragments of 110–120 kDa, 75 kDa, and 37 kDa (Ruoslahti 1981;<br />

Hayashi and Yamada 1983; Zardi et al. 1985; Nagai et al. 1991) derived from the central<br />

portion of the protein. Such fragments of fibronectin retained similar cell adhesive activities as<br />

those of intact fibronectin. The cell adhesive activity attributed to these fragments was initially<br />

localized to the tenth type III module in the form of an 11.5 kDa pepsin fragment (Pierschbacher,<br />

Heparin Collagen<br />

ED-B Cell<br />

Heparin Fibrin<br />

ED-A III-C-S<br />

H2NI II III II II II<br />

I<br />

II<br />

I<br />

II<br />

I<br />

II<br />

I<br />

II II<br />

I I<br />

III<br />

II II II II II II<br />

I I I I I<br />

PHSRN RGD<br />

II<br />

I<br />

II<br />

I<br />

II<br />

I<br />

LDV<br />

II<br />

I<br />

REDV<br />

I I COOH<br />

S S<br />

S S<br />

FIGURE 32.4 Structure of fibronectin. Fibronectin is composed of three types of internal repeating modules<br />

designated as Type I, Type II, and Type II. The ED-A, ED-B, and III-C-S modeules can be present or absent in<br />

some forms of fibronectin as a result of alternative splicing. The binding domains are indicated at the top. The<br />

central cell-binding domain consists of ninth and tenth Type III nodules, containing the minimal PHSRN and<br />

RGD cell recognition sequence (Adapted from Mosher, D. F., Fibronectin, Academic Press, NY, p.474, 1989.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 651<br />

Hayman, and Ruoslahti 1981) and to a smaller peptide with the sequence Gly-Arg-Gly-Asp-Ser<br />

(GRGDS). Although the 11.5 kDa fragment and synthetic peptides containing the RGD sequence<br />

can inhibit fibronectin cell-adhesive functions in vitro and in vivo when added as soluble inhibitors<br />

(Pierschbacher and Ruoslahti 1984; Lash, Linask, and Yamada 1987), they only poorly promote<br />

cell adhesion mediated <strong>by</strong> the major fibronectin receptor a5b1 integrin, and their affinities are too<br />

low to be estimated in direct binding studies (Akiyama and Yamada 1985), suggesting that<br />

sequences outside of the tenth type III module are also important for maximal cell binding<br />

and adhesion.<br />

This additional fibronectin sequence that is important for maximal cell adhesive activity was<br />

identified and characterized using a series of mutants of the central fibronectin cell adhesive region<br />

expressed in E. coli (Aota, Nagai, and Yamada 1991) and anti-fibronectin monoclonal antibodies<br />

(Nagai et al. 1991). Fragments containing the RGD sequence but truncated approximately 10–<br />

14 kDa to the amino-terminal side of the RGD sequence were !4% as active as compared to intact<br />

fibronectin in mediating cell adhesion, whereas fragments containing these amino terminal<br />

sequences retained O97% of the activity of intact fibronectin. The novel, amino-terminal (non-<br />

RGD) site appeared to act synergistically with the RGD site to promote better cell adhesion, leading<br />

to its designation as a synergistic adhesive site or synergy site.<br />

The biological function of the synergistic cell adhesive site was characterized <strong>by</strong> using a panel<br />

of anti-fibronectin monoclonal antibodies (mAbs) developed to bind a 37 kDa cell adhesive fibronectin<br />

fragment. One of these antibodies, designated 8E3, bound to the ninth type III module at a<br />

site approximately 14–16 kDa to the amino-terminal side of the RGD sequence and close to the<br />

synergy site identified <strong>by</strong> mutational studies. Other mAbs that inhibited cell adhesion such as 333<br />

(Akiyama et al. 1985) and 16G3 bound to the tenth type III module and inhibited the RGD site.<br />

Interestingly, there was also an antibody, designated 13G12, that bound to fibronectin between the<br />

RGD and synergy sites but did not inhibit cell adhesion. Antibodies that bound near the RGD and<br />

synergy sites could each individually inhibit cell spreading at high concentrations. Furthermore,<br />

mAb 8E3 and mAb 16G3, at concentrations too low to inhibit individually spreading, can be highly<br />

inhibitory in combination, underscoring the synergistic nature of the two cell adhesive sites.<br />

Antibody inhibition experiments have also shown that both the RGD and synergy sites function<br />

in cell migration and cytoskeleton assembly on fibronectin substrates as well as in the assembly of a<br />

fibronectin extracellular matrix (Nagai et al. 1991), indicating that both sites are required for a<br />

range of fibronectin activities.<br />

32.3.3 RGD-MEDIATED METASTASIS INHIBITION<br />

A major cause of morbidity and death as a result of cancer is the metastasis of cells from the primary<br />

tumor to distant sites where secondary tumors become established. As schematically shown in<br />

Figure 32.5, metastasis is a multistep process. These steps include detachment of cells from the<br />

tumor mass, degradation of basement membrane, migration to and invasion into the vascular or<br />

lymphatic systems, arrest at a distant site, adhesion to the vascular endothelium, degradation of<br />

basement membrane, extravasation, and migration to and proliferation at the secondary site. Many<br />

of these steps require cell adhesive interactions or loss of adhesion. Tumor cell adhesion to<br />

components of the extracellular matrix and basement membranes is mediated <strong>by</strong> specific cell<br />

surface receptors that bind to extracellular adhesive proteins. The fibronectin-integrin system has<br />

provided a valuable model system for the study of molecular mechanisms of ligand-receptor<br />

interactions involved in cell adhesive steps in metastasis.<br />

Several different assay systems have been used to directly analyze the role of fibronectin and<br />

integrins in in vivo metastasis, but they all fall into two broad classes. The experimental metastasis<br />

models mainly involve the intravenous injection of tumor cell suspensions into mice and subsequent<br />

quantitation of either the number of metastatic colonies or the sizes of colonies usually<br />

in the lungs or liver. Experimental metastasis models, however, exclude the earlier steps of the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


652<br />

Basement<br />

Membrane<br />

Carcinoma<br />

Invasion<br />

Sarcoma<br />

Capillary<br />

Migration<br />

Nanotechnology for Cancer Therapy<br />

Blood vessel or<br />

Lymphatic system<br />

Proliferation<br />

Extravasation<br />

Arrest<br />

FIGURE 32.5 Major steps in metastasis. Tumor cells are shown invading the circulatory system from a<br />

carcinoma <strong>by</strong> degrading and migrating through a basement membrane or from a sarcoma. Cells separate<br />

from the primary tumor, enter the vasculature, and eventually arrest in capillaries where they penetrate and<br />

migrate through basement membrane and underlying connective tissue to the metastatic site where colonization<br />

and proliferation occur.<br />

metastatic process involving detachment from the tumor mass and invasion into the vasculature<br />

and, subsequently, replicate only following the final steps of hematogenous metastasis. Spontaneous<br />

metastasis models where tumors or cells are implanted into mice and subsequent<br />

metastases to distant organs are quantitated either <strong>by</strong> counting colonies or measuring colony size<br />

are more complicated, but they are required to analyze of the earlier steps of the metastatic cascade.<br />

Synthetic peptides containing the Gly-Arg-Gly-Asp-Ser (GRGDS) sequence derived from<br />

fibronectin were assayed <strong>by</strong> examining their effects on lung colonization of B16-F10 murine<br />

melanoma cells in syngeneic C57BL/6 mice using an experimental metastasis model (Humphries,<br />

Olden, and Yamada 1986). GRGDS peptide that had a circulatory half-life of approximately eight<br />

minutes specifically inhibited lung colonization in a concentration-dependent manner. The major<br />

effect of the GRGDS peptide appeared to be to inhibit arrest of melanoma cells in the lungs without<br />

affecting the size of either melanoma cell clusters in suspension or lung colonies. The efficacy of<br />

peptide treatment was unaltered in animals with impaired platelet function and in animals lacking<br />

natural killer cells. Taken together, these results suggest that the GRGDS peptide could inhibit<br />

metastasis <strong>by</strong> disrupting an early adhesive process. As shown in Figure 32.6, a single administration<br />

of the GRGDS peptide also had the dramatic effect of increasing the survival of mice when<br />

co-injected with melanoma cells (Humphries, Yamada, and Olden 1988).<br />

RGD peptides and RGD mimetics are very promising molecules. RGD peptides not only inhibit<br />

metastatic colony formation but a single administration of the peptide can dramatically increase<br />

animal survival also however, the use of such agents have some drawbacks. One potential problem is<br />

the relatively high concentrations of peptide required for activity. Possible solutions to the problems<br />

may include the use of cyclic peptides and designer peptides that reportedly increased the affinity for<br />

cell surface integrins or the use of repeating polymers containing RGD units (Saiki et al. 1989;<br />

Murata et al. 1991; Yamamoto et al. 1994). Polymers containing RGD sequences have the dual<br />

advantages of augmenting the affinity of small peptide sequences through multivalent interactions<br />

and the possibility of solving the problem of the relatively short circulatory half-life of GRGDS<br />

peptides. An alternative approach would be to develop novel, small, synthetic integrin antagonists.<br />

Such compounds that inhibit RGD-dependent and LDV-dependent cell adhesion have already been<br />

synthesized and tested in mice (Greenspoon et al. 1993, 1994).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 653<br />

Survival (%)<br />

100<br />

75<br />

50<br />

25<br />

0<br />

•<br />

•<br />

•<br />

Treated<br />

•<br />

•<br />

•<br />

•<br />

Untreated<br />

•<br />

•<br />

30 60 120 180 240 300 360<br />

Days post injection<br />

FIGURE 32.6 Effect of GRGDS peptide on survival of C57BL/6 mice injected with B16F10 murine melanoma<br />

cells. Mice were co-injected with 3!10 4 B16F10 melanoma cells mixed with 3 mg GRGDS paptide.<br />

(Data are from Humphries et al., J. Clin. Invest., 81, 782, 1988.)<br />

32.4 RGD-MODIFIED LIPOSOMES FOR CANCER THERAPEUTICS<br />

Cytotoxic drug incorporation into liposomes has been reported since the mid-1970s. These early<br />

reports laid the groundwork for selecting therapeutic agents amenable to incorporation <strong>by</strong> liposomes<br />

as well as determining optimal liposome size and net charge required for effective drug<br />

delivery (Gregoriadis 1976). Early reports also noted the problems associated with liposomemediated<br />

drug delivery; liposomes were removed from circulation <strong>by</strong> fixed macrophages in the<br />

RES, particularly in the liver and spleen. Although first-generation liposome-incorporated drugs<br />

may be effective in macrophage-related diseases, particularly in the liver and spleen, many other<br />

tumors would require other drug-targeting mechanisms. Later studies refined many of the considerations<br />

required for effective liposome-mediated drug delivery (Drummond et al. 1999).<br />

In 2005, Pattillo et al. developed RGD-modified immunoliposmes for targeting the antivascular<br />

drug combretastatin to irradiated mouse melanomas. Combretastatin was incorporated into liposomes<br />

with surfaces modified <strong>by</strong> the addition of cyclo(Arg-Gly-Asp-D-Phe-Cys) (RGD) to create<br />

an immunoliposome. In this transplanted tumor model (B16-F10 melanoma), there was no significant<br />

increase in the tumor volume when treated with a single dose of 5-Gy radiation and RGD<br />

modified immunoliposomes (14.5 mg/kg of combretastatin) during the initial six days post treatment;<br />

all other treatment groups exhibited exponential growth curves after three days. The<br />

treatment resulted in a 5.1-day tumor growth delay compared to untreated controls. These findings<br />

indicate that preferential targeting of antivascular drugs to irradiated tumors results in significant<br />

tumor growth delay.<br />

Various strategies are adopted to target the sites of tumor-associated angiogenesis or other<br />

inflammatory environments with liposome-immobilized ligands where adhesion receptors (or<br />

ligands) serve as molecular targets (Figure 32.7). Targeting can be achieved using RGD<br />

or YIGSR immobilized on liposome surface and interacting with integrin receptors on normal or<br />

malignant cells that over-express them. Targeting can be achieved using synthetic antibodies<br />

anchored on the liposomes and bind integrins and other cell adhesion molecules molecules <strong>by</strong><br />

mimicking their natural ligands. Targeting can be achieved <strong>by</strong> immobilizing specific carbohydrate<br />

ligands (sLe x and sLe a ) on liposomes that show specific interaction with selectin. These strategies<br />

are directed either to block the normal recruitment of leucocytes during vasculogenesis or for<br />

selective targeting of chemotherapeutic agents to tumor-associated vasculature.<br />

Cyclic RGD peptide, cyclo(Arg-Gly-Asp-Phy-Lys) anchored sterically stabilized liposomes<br />

(RGD-SL) were investigated for selective and preferential presentation of carrier contents at angiogenic<br />

endothelial cells over-expressing alphavbeta3 integrins on and around tumor tissue and for<br />

assessing their targetabilty (Dubey et al. 2004; Dubey 2005). Liposomes were prepared using<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


654<br />

YIGSR<br />

RGD<br />

distearoylphosphatidylcholine (DSPC), cholesterol, and distearoylphosphatidylethanolamine-polyethyleneglycol-RGD<br />

peptide conjugate (DSPE-PEG-RGD) in a molar ratio 56:39:5. The control<br />

RAD peptide anchored sterically stabilized liposomes (RAD-SL), and liposome with 5 mol% PEG<br />

(SL) without peptide conjugate that had similar lipid composition were used for comparison. The<br />

average size of all liposome preparations prepared was approximately 105 nm and maximum drug<br />

entrapment was 10.5C/K1.1%. In vitro endothelial cell binding of liposomes exhibited 7-fold<br />

higher binding of RGD-SL to HUVEC in comparison to the SL and RAD-SL (Figure 32.8a).<br />

Table 32.4 and Figure 32.9 indicate RGD-modified liposomes mediated spontaneous lung<br />

metastasis and angiogenesis assays, respectively. RGD peptide-anchored liposomes are significantly<br />

(p!0.01) effective in the prevention of lung metastasis and angiogenesis compared to<br />

sLe x<br />

sLe a<br />

(a) (b)<br />

FIGURE 32.7 Liposome-immobilized ligands to target the cell adhesion molecules on tumor vasculature and<br />

extracellular matrix. These liposomes serve as artificial leucocyte to mediate interactions with integrins (a),<br />

integrins and other CAM molecules (b), and selectins (c) with the help of immobilized ligands or anti-receptor<br />

antibody. (Adapted from Vyas, S. P., and Khar, R. K., Novel Carrier Systems, CBS publishers, New Delhi,<br />

p.535, 2001.)<br />

% of maximum HUVEC binding<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0 0<br />

SL RAD-SL RGD-SL RGD-SL + RGD-SL +<br />

Control Free 5-FU SL RGD-SL RAD-SL<br />

(a)<br />

free RGD<br />

Formulation<br />

free RAD<br />

(b)<br />

Formulation<br />

FIGURE 32.8 (a) Binding of RGD-modified liposomes with HUVEC cells. (b) Effect of various formulations<br />

on tumor angiogenesis <strong>by</strong> an intradermal injection of B16F10 melanoma. Results are given as means (S.D.)<br />

NZ5(P!0.01). 5-FU, SL, RGD-SL, and RAD-SL indicate free 5-fluorouracil, stealth liposomes (SL), RGDanchored<br />

stealth liposomes (RGD-SL), and RAD-anchored stealth liposomes (RAD-SL), respectively<br />

(Adapted from Dubey, P. K., et al., J. Drug. Target., 12, 257, 2004).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Angiogenesis (No. of vessels)<br />

(c)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

Nanotechnology for Cancer Therapy


RGD-Modified Liposomes for Tumor Targeting 655<br />

TABLE 32.4<br />

Effect of Free 5-FU, SL, RGD-SL, and RAD-SL on Spontaneous Lung Metastasis in BALB/c<br />

Mice <strong>by</strong> an Intra-Footpad Injection of B16F10 Melanoma<br />

Formulations<br />

Average Number<br />

of Metastasis/<br />

Mouse<br />

Average Number of Metastasis/<br />

Tumor Bearing Mouse<br />

Number of Metastasis Free Mice<br />

(NZ10)<br />

Control 52G15.5 52G15.5 0<br />

Free 5-FU 46G13.6 46G13.6 0<br />

SL 26G8.6 32G7.65 2<br />

RGD-SL 8G4.2* 16G6.58 5<br />

RAD-SL 28G6.5 31G8.7 1<br />

Source: Dubey et al., J. Drug. Target.,12, 2004. *P!0.01<br />

free 5-FU, SL, and RAD-SL. In therapeutic experiments, 5-FU, SL, RGD-SL, and RAD-SL were<br />

intravenously administered on day 4 at the dose of 10 mg 5-FU/kg body weight to B16F10 tumor<br />

bearing BALB/c mice, resulting in effective regression of tumors compared with free 5-FU, SL, and<br />

RAD-SL (Dubey et al. 2004). Results indicated that cyclic RGD peptide anchored sterically<br />

stabilized liposomes bearing 5-FU were more significantly (p!0.01) active against primary<br />

tumor and metastasis than the non-targeted sterically stabilized liposomes and free drug<br />

(Figure 32.9). Vascular damage and reduced vascularization were recorded in tumors treated<br />

with RGD-SL, but not in the tumor treated with SL (Figure 32.10). These results also confirmed<br />

that RGD peptide-anchored liposomes cause a marked improvement in therapeutic efficacy and<br />

growth inhibitory effect on tumors.<br />

RGD-modified sterically stabilized liposomes have also been evaluated to improve the antitumor<br />

efficacy of doxorubicin (Schiffelers et al. 2003; Xiong et al. 2005). RGD peptide was coupled<br />

to the distal end of the poly (ethylene glycol)-coated liposomes. RGD peptide anchored sterically<br />

Tumor volume (mm 3 )<br />

2000<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0<br />

4 6 8 10 12 14 16 18 20<br />

Days following tumor implantation<br />

FIGURE 32.9 Anti-tumor efficacy of control (!), free 5-FU (%), SL (&), RAD-SL (C), and RGD-SL (:)in<br />

mice bearing B16F10 melanoma subcutaneously inoculated in the back. The arrow represents the time of<br />

injection. Results are given as means (S.D.) NZ10 (P!0.01). (Adapted from Dubey, P.K. et al., J. Drug.<br />

Target., 12, 257, 2004.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


656<br />

FIGURE 32.10 (See color insert following page 522.) Histopathological analysis (hematoxylin and eosin stain)<br />

of B16F10 tumors treated with (a) Stealth liposomes and (b) RGD-modified stealth liposomes (250X). (Adapted<br />

from Dubey, P. K., Development and characterization of biochemical based strategies for tumor targeting. Ph. D.<br />

thesis, Department of Pharmaceutical Sciences, Dr. H. S. Gour University, Sagar (MP) India, 2005.)<br />

stabilized liposomes (dose 5 mg doxorubicin/kg) have demonstrated prolonged circulation time and<br />

increased tumor accumulation and effective retardation in tumor growth compared with sterically<br />

stabilized liposomes without RGD modification.<br />

Holig and coworkers have isolated from phage display RGD motif libraries with novel highaffinity<br />

cyclic RGD peptides on the basis of their selectivity towards endothelial or melanoma cells<br />

(Holig et al. 2004). Although the starting sequences contained only two cysteine residues flanking<br />

the RGD motif, several of the isolated peptides possessed four cysteine residues. A high-affinity<br />

peptide (RGD10) (DGARYCRGDCFDG) constrained <strong>by</strong> only one disulfide bond was used to<br />

generate novel lipopeptides composed of a lipid anchor, a short flexible spacer, and the peptide<br />

ligand conjugated to the spacer end. Incorporation of RGD10 lipopeptides into liposomes resulted<br />

in specific and efficient binding of the liposomes to integrin-expressing cells. In vivo experiments<br />

applying doxorubicin-loaded RGD10 liposomes in a C26 colon carcinoma mouse model demonstrated<br />

improved efficacy compared with free doxorubicin and untargeted liposomes.<br />

Administration of large amounts of synthetic peptides based on the Arg-Gly-Asp (RGD)<br />

sequence has been shown to suppress tumor metastasis. To overcome the rapid degradation of<br />

peptides in the circulation, an RGD mimetic, L-arginyl-6-aminohexanoic acid (NOK), was synthesized<br />

and conjugated with phosphatidylethanolamine (PE) (NOK-PE) for liposomalization<br />

(Kurohane, Namba, and Oku 2000). Cell adhesion assays revealed that B16BL6 murine melanoma<br />

cells adhered to immobilized NOK-PE. This adhesion was inhibited <strong>by</strong> the addition of either<br />

soluble RGDS or NOK at similar concentrations in a dose-dependent manner. Administration of<br />

NOK-PE liposomes (equivalent to approximately 500 microg RGD peptides) via the tail vein<br />

completely inhibited lung colonization of B 16BL6 cells. The same dose of soluble NOK was<br />

not effective in inhibition of the tumor metastasis. In addition, injection of NOK-PE liposomes via<br />

the tail vein inhibited spontaneous lung metastasis of B16BL6 cells from the primary tumor site in<br />

the hind footpad. These results suggest that NOK, a structural mimetic of RGD, is capable of<br />

suppressing metastasis <strong>by</strong> blocking the binding of the integrins present on tumor cells to the RGDcontaining<br />

extracellular matrix.<br />

32.5 RGD-MODIFIED LIPOSOMES IN CANCER GENE THERAPY<br />

Nanotechnology for Cancer Therapy<br />

Gene therapy, <strong>by</strong> definition, aims at modifying the genetic program of a cell toward a therapeutic or<br />

prophylactic goal. Several gene therapy strategies for tumors are currently under evaluation and<br />

some of them include: the modification of the function of oncogenes and tumor suppressor genes;<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 657<br />

the modification of the host immune response toward the tumor; the disruption of the tumor<br />

neovascularization; the lysis of tumor cells with replication-competent viruses, and suicide gene<br />

therapy where an inactive prodrug is converted into a cytotoxic drug <strong>by</strong> gene-expressed enzymes.<br />

Both viral and non-viral gene vectors have been investigated and well documented (Ledley 1995).<br />

However, the inefficiency of current gene vectors in infecting targeted cells and their inability to<br />

selectively access the diseased cells systemically distributed are two major impediments that have<br />

to be overcome for further successful clinical applications.<br />

RGD peptides have been used to taget the lipid-protamine-DNA (LPD) lipopolyplexes to tumor<br />

cells (MDA-MB-231), expressing appropriate integrin receptors (Harvie et al. 2003). The incorporation<br />

of pegylated lipid into Lipid-Protamine-DNA (LPD-PEG) lipopolyplexes causes a<br />

decrease of their in vitro transfection activity. This can be partially attributed to a reduction in<br />

particle binding to cells. To restore particle binding and specifically target LPD formulations to<br />

tumor cells, the lipid-peptide conjugate DSPE-PEG5K-succinyl-ACDCRGDCFCG-COOH<br />

(DSPE-PEG5K-RGD-4C) was generated and incorporated into LPD formulations (LPD-PEG-<br />

RGD). LPD-PEG-RGD was characterized with respect to its biophysical and biological properties.<br />

The incorporation of DSPE-PEG5K-RGD-4C ligands into LPD formulations results in a 5- and<br />

15-fold increase in the LPD-PEG-RGD binding and uptake, respectively, over an LPD-PEG formulation.<br />

Enhancement of binding and uptake resulted in a 100-fold enhancement of transfection<br />

activity. Moreover, this transfection enhancement was specific to cells expressing appropriate<br />

integrin receptors (MDA-MB-231). Huh7 cells, known for their low level of avb 3 and a vb 5 integrin<br />

expression, failed to show RGD-mediated transfection enhancement. This transfection enhancement<br />

can be abolished in a competitive manner using free RGD peptide, but not an RGE control<br />

peptide. Results demonstrated RGD-mediated enhanced LPD-PEG cell binding and transfection in<br />

cells expressing the integrin receptor. These formulations provide the basis for effective, targeted,<br />

systemic gene delivery.<br />

In 2002, Fahr et al. developed a novel liposomal vector (Artificial Virus Particles; AVPs) for<br />

cancer gene therapy. Artificial virus-like particles (AVPs) represent a novel type of liposomal<br />

vector, resembling retroviral envelopes. AVPs are serum-resistant and non-toxic and can be<br />

endowed with a peptide ligand as a targeting device. These workers investigated if AVPs carrying<br />

cyclic peptides with an RGD integrin-binding motif (RGD-AVPs) were suitable for the specific and<br />

efficient transduction of human melanoma cells. In the experimentation, they prepared Plasmid<br />

DNA-low molecular weight non-linear polyethyleneimine complexes and packaged into anionic<br />

liposomes. Transduction efficiencies were determined after transient transfection of different cell<br />

lines in serum-free medium using green fluorescent protein or luciferase reporter genes. It was<br />

found that RGD-AVPs transduced human melanoma cells with high efficiencies of O60%. Efficient<br />

transduction was clearly dependent on the presence of the cyclic RGD ligand and was<br />

selective for melanoma cells. Using the melanocyte-specific tyrosinase promoter to drive transgene<br />

expression, the specificity of the vector system could be further enhanced.<br />

REFERENCES<br />

Akiyama, S. K. and Yamada, K. M., Synthetic peptides competitively inhibit both direct binding to fibroblasts<br />

and functional biological assays for the purified cell-binding domain of fibronectin, J. Biol. Chem.,<br />

260, 10402, 1985.<br />

Akiyama, S. K. et al., The interaction of fibronectin fragments with fibroblastic cells, J. Biol. Chem., 260,<br />

13256, 1985.<br />

Aota, S., Nagai, T., and Yamada, K. M., Characterization of regions of fibronectin besides the arginineglycine-aspartic<br />

acid sequence required for adhesive function of the cell-binding domain using sitedirected<br />

mutagenesis, J. Biol. Chem., 266, 15938, 1991.<br />

Arap, W., Pasqualini, R., and Ruoslahti, E., Cancer treatment <strong>by</strong> targeted drug delivery to tumor vasculature in<br />

a mouse model, Science, 279, 377, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


658<br />

Nanotechnology for Cancer Therapy<br />

Arora, N. et al., Vascular endothelial growth factor chimeric toxin is highly active against endothelial cells,<br />

Cancer Res., 59, 183, 1999.<br />

Boehm, T. et al., Antiangiogenic therapy of experimental cancer does not induce acquired drug resistance,<br />

Nature, 390, 404, 1997.<br />

Bowditch, R. D. et al., Identification of a novel integrin-binding site in fibronectin: differential utilization <strong>by</strong> b,<br />

integrins, J. Biol. Chem., 269, 10856, 1994.<br />

Brooks, P. C., Clark, R. A., and Cheresh, D. A., Requirement of vascular integrin avb3 for angiogenesis,<br />

Science, 264, 569, 1994a.<br />

Brooks, P. C. et al., Integrin Alpha-v Beta-3 antagonists promote tumor regression <strong>by</strong> inducing apoptosis of<br />

angiogenic blood vessels, Cell, 79, 1157, 1994b.<br />

Brown, P. M. and Silvius, J. R., Mechanisms of delivery of liposome-encapsulated cytosine arabinoside to<br />

CV-1 cells in vitro: Fluorescence-microscopic and cytotoxicity studies, Biochim. Biophys. Acta, 1023,<br />

341, 1990.<br />

Burrows, F. J. et al., Endoglin is an endothelial cell proliferation marker that is selectively expressed in tumor<br />

vasculature, Clin. Cancer Res., 1, 1623, 1995.<br />

Carron, C. P. et al., A. peptidomimetic antagonist of the intergin alpha (V) beta 3 inhibits Leydig cell tunour<br />

growth and the development of hypercalcemia of malignancy, Cancer Res., 58, 1930, 1998.<br />

Carsons, S. E., Fibronectin in Health and Disease, CRC Press, Boca Raton, 1990.<br />

Chang, Z., Choon, A., and Friedl, A., Endostatin binds to blood vessels in situ independent of heparan sulfate<br />

and does not compete for fibroblast growth factor-2 binding, Am. J. Pathol., 155, 71, 1999.<br />

Curley, G. P., Blum, H., and Humphries, M. J., Integrin antagonists, Cell. Mol. Life Sci., 56, 427, 1999.<br />

Curnis, F. et al., Enhancement of tumor neaosis factor alpha antitumour immunothecapetic properties <strong>by</strong><br />

targeted delivery to aminopeptidase N (CD13), Nat. Biotechnol., 18, 1185, 2000.<br />

Davol, P. et al., Saporin toxins directed to basic fibroblast growth factor receptors effectively target human<br />

ovarian teratocarcinoma in an animal model, Cancer, 76, 79, 1995.<br />

Denekamp, J., Angiogenesis, neovascular proliferation and vascular pathophysiology as targets for cancer<br />

therapy, Br. J. Radiol., 66, 181, 1993.<br />

Drummond, D. C. et al., Optimizing liposomes for delivery of chemotherapeutic agents to solid tumors,<br />

Pharmacol. Rev., 51, 691, 1999.<br />

Dubey, P. K., Development and characterization of biochemical based strategies for tumor targeting. Ph. D.<br />

thesis, Department of Pharmaceutical Sciences, Dr. H. S. Gour University, Sagar (MP) India, 2005.<br />

Dubey, P. K. et al., Liposomes modified with cyclic RGD peptide for tumor targeting, J. Drug Target, 12, 257,<br />

2004.<br />

Dvorak, H. F. et al., Distribution of vascular permeability factor (vascular endothelial growth factor) in tumors:<br />

Concentration in tumor blood vessels, J. Exp. Med., 174, 1275, 1991.<br />

Eggermont, A. M. et al., Isolated limb perfusion with high-dose tumor necrosis factor-alpha in combination<br />

with interferon-gamma and melphalan for nonresectable extremity soft tissue sarcomas: A multicenter<br />

trial, J. Clin. Oncol., 14, 2653, 1996.<br />

Eller<strong>by</strong>, H. M. et al., Anticancer activity of targeted pro-apoptotic peptides, Natl. Med., 5, 1032, 1999.<br />

Fahr, A. et al., A new colloidal lipidic system for gene therapy, J. Liposome Res., 12, 37, 2002.<br />

Folkman, J., Angiogenesis and angiogenesis inhibition: An overview, EXS, 79, 1, 1997.<br />

Forssen, E. and Willis, W., Ligand targeted liposomes, Adv. Drug Del. Rev., 29, 249, 1998.<br />

Gabizon, A. and Papahadjopoulos, D., The role of surface charge and hydrophilic groups on liposome clearance<br />

in vivo, Biochim. Biophys. Acta, 1103, 94, 1992.<br />

Gerlag, D. M. et al., Suppression of murine collagen-induced arthritis <strong>by</strong> targeted apoptosis of synovial<br />

neovasculature, Arthritis Res., 3, 357, 2001.<br />

Giblin, P. A. and Kelly, T. A., Antagonists of b2 integrin-mediated cell adhesion, Ann. Rep. Med. Chem., 36,<br />

181, 2001.<br />

Greenspoon, N. et al., Structural analysis of integrin recognition and the inhibition of integrin-mediated cell<br />

function <strong>by</strong> novel nonpeptidic surrogates of the Arg-Gly-Asp sequence, Biochemistry, 32, 1001, 1993.<br />

Greenspoon, N. et al., Novel 0 F-S-CH, peptide-bond replacement and its utilization in the synthesis of<br />

nonpeptidic surrogates of the Leu-Asp-Val sequence that exhibit specific inhibitory activities on<br />

CD4 C T cell binding to fibronectin, Int. J. Pept. Res., 43, 417, 1994.<br />

Gregoriadis, G., The carrier potential of liposomes in biology and medicine (first of two parts), N. Engl.<br />

J. Med., 295, 704, 1976.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 659<br />

Griffioen, A. W. et al., CD44 is an activation antigen on human endothelial cells, involvement in tumor<br />

angiogenesis, Blood, 90, 1150, 1997.<br />

Gyongyossy-Issa, M. I. C., Muller, W., and Devine, D., The covalent coupling of Arg-Gly-Asp-containing<br />

peptides to liposomes: Purification and biochemical function of lipopeptide, Arch. Biochem. Biophys.,<br />

353, 101, 1998.<br />

Hagemeier, H. H. et al., A monoclonal antibody reacting with endothelial cells of budding vessels in tumors<br />

and inflammatory tissues, and non-reactive with normal adult tissues, Int. J. Cancer, 38, 481, 1986.<br />

Harvie, P. et al., Targeting of lipid-protamine-DNA (LPD) lipopolyplexes using RGD motifs, J. Liposome<br />

Res., 13, 231, 2003.<br />

Hayashi, M. and Yamada, K. M., Domain structure of the carboxy-terminal half of human plasma fibronectin,<br />

J. Biol. Chem., 258, 3332, 1983.<br />

Holig, P. et al., Novel RGD lipopeptides for the targeting of liposomes to integrin-expressing endothelial and<br />

melanoma cells, Protein Eng. Des. Sel., 17, 433, 2004.<br />

Hood, J. D. et al., Tumor regression <strong>by</strong> targeted gene delivery to the neovasculature, Science, 269, 2404, 2002.<br />

Huang, X. et al., Tumor infarction in mice <strong>by</strong> antibody-directed targeting of tissue factor to tumor vasculature,<br />

Science, 275, 547, 1997.<br />

Humphries, M. J., Olden, K., and Yamada, K. M., A synthetic peptide from fibronectin inhibits experimental<br />

metastasis of murine melanoma cells, Science, 233, 466, 1986.<br />

Humphries, M. J., Yamada, K. M., and Olden, K., Investigation of the biological effects of anti-cell adhesive<br />

synthetic peptides that inhibit experimental metastasis of B16-F10 murine melanoma cells, J. Clin.<br />

Investig., 81, 782, 1988.<br />

Hynes, R. O., Fibronectins, Springer-Verlag, NY, p. 546, 1990.<br />

Hynes, R. O., Integrins: Versatility, modulation, and signaling in cell adhesion, Cell, 69, 11, 1992.<br />

Kennel, S. J. and Mirzadeh, S., Vascular targeted radioimmunotherapy with 2I3 Bi-an-particle emitter, Nucl.<br />

Med. Biol., 25, 241, 1998.<br />

Kirpotin, D. et al., Sterically stabilized anti-HER2 liposomes: Design and targeting to human breast cancer<br />

cells in vitro, Biochemistry, 36, 66, 1997.<br />

Koivunen, E. et al., Tumor targeting with a selective gelatinase inhibitor, Natl. Biotechnol., 17, 768, 1999.<br />

Kurohane, K., Namba, Y., and Oku, N., Liposomes modified with a synthetic Arg-Gly-Asp mimetic inhibit<br />

lung metastasis of B16BL6 melanoma cells, Life Sci., 68, 273, 2000.<br />

Lash, J. W., Linask, K. K., and Yamada, K. M., Synthetic peptides that mimic the adhesive recognition signal<br />

of fibronectin: Differential effects on cell–cell and cell–substratum adhesionin embryonic chick cells,<br />

Dev. Biol., 123, 411, 1987.<br />

Ledley, F. D., Nonviral gene therapy: the promise of genes as pharmaceutical products, Hum. Gene. Ther., 6,<br />

1129,1995.<br />

Lopes de Menezes, D. E., Pilarski, L. M., and Allen, T. M., In vitro and in vivo targeting of immunoliposomal<br />

doxorubicin to human B-cell lymphoma, Cancer Res., 58, 3320, 1998.<br />

Lundberg, B., Hong, K., and Papahadjopoulos, D., Conjugation of apolipoprotein B with liposome and<br />

targeting of cells in culture, Biochim. Biophys. Acta, 1149, 305, 1993.<br />

Martin, F. J., Clinical pharmacology and antitumor efficacy of DOXIL (pegylated liposomal doxorubicin), In<br />

Medical Applications of Liposomes, Lasic, D. D. and Papahadjopoulos, D., Eds., Elsevier Science BV,<br />

New York, p. 635, 1998.<br />

Maruyama, K. et al., Characterization of in vitro immunoliposome targeting to pulmonary endothelium,<br />

J. Pharm. Sci., 79, 978, 1990.<br />

Miller, D.W. et al., Elevated expression of endoglin, a component of the TGF-b-receptor complex, correlates<br />

with proliferation of tumor endothelial cells, Int. J. Cancer, 81,568,1999.<br />

Moser, T. L. et al., Angiostatin binds ATP synthase on the surface of human endothelial cells, Proc. Natl Acad.<br />

Sci. USA, 96, 2811, 1999.<br />

Mosher, D. F., Fibronectin, Academic Press, New York, 1989.<br />

Murata, J. et al., Molecular properties of poly(RGD) and its binding capacities to metastatic melanoma cells,<br />

Int. J. Pept. Prot. Res., 38, 212, 1991.<br />

Nishiya, T. and Sloan, S., Interaction of RGD liposomes with platelets, Biochem. Biophys. Res. Commun., 224,<br />

242, 1996.<br />

Nguyen, M. et al., A role for sialyl Lewis-X/A glycoconjugates in capillary morphogenesis, Nature, 365,<br />

267,1993. Erratum in: Nature, 366, 368, 1993.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


660<br />

Nanotechnology for Cancer Therapy<br />

Niederman, T.M. et al., Antitumor activity of cytoxic T lymphocytes engineered to target vascular endothelial<br />

growth factor receptors, Proc. Natl. Acad. Sci. USA, 99, 709, 2002.<br />

Obara, M. and Yoshizato, K., Possible involvement of the interaction of the a, subunit of the a, b, integrin with<br />

the synergistic region of the central cell-binding domain of fibronectin in cells to fibronectin binding,<br />

Exp. Cell Res., 216, 273, 1995.<br />

Olson, T. A. et al., Targeting the tumor vasculature: inhibition of tumor growth <strong>by</strong> a vascular endothelial<br />

growth factor toxin conjugate, Int. J. Cancer, 73, 865, 1997.<br />

Papahadjopoulos, D. and Gabizon, A. A., Sterically stabilized (stealth) liposomes: Pharmacological properties<br />

and drug carrying potentialin cancer, In Liposomes as Tools in Basic Research and Industry, Philippot,<br />

J. R. and Schuber, F., Eds., CRC Press, Boca Raton, p. 177, 1995.<br />

Papahadjopoulos, D. et al., Sterically stabilized liposomes: Improvements in pharmacokinetics and antitumor<br />

therapeutic efficacy, Proc. Natl Acad. Sci. USA, 88, 11460, 1991a.<br />

Papahadjopoulos, D. et al., Sterically stabilized liposomes: Improvements in pharmacokinetics and antitumor<br />

therapeutic efficacy, Proc. Natl Acad. Sci. USA, 88, 11460, 1991b.<br />

Park, J. W. et al., Development of anti-p185 HER2 immunoliposomes for cancer therapy, Proc. Natl Acad. Sci.<br />

USA, 92, 1327, 1995.<br />

Pattillo, C. B. et al., Targeting of the antivascular drug combretastatin to irradiated tumors results in tumor<br />

growth delay, Pharm. Res., 22, 1117, 2005.<br />

Pierschbacher, M. D. and Ruoslahti, E., Cell attachment activity of fibronectin can be duplicated <strong>by</strong> small<br />

synthetic fragments of the molecule, Nature, 309, 30, 1984.<br />

Pierschbacher, M. D. and Ruoslahti, E., Influence of stereochemistry of the sequence Arg-Gly- Asp-Xaa on<br />

binding specificity in cell adhesion, J. Biol. Chem., 262, 17294, 1987.<br />

Pierschbacher, M. D., Hayman, E. G., and Ruoslahti, E., Location of the cell attachment site in fibronectin with<br />

monoclonal antibodies and proteolytic fragments of the molecule, Cell, 26, 259, 1981.<br />

Ramakrishnan, S. et al., Vascular endothelial growth factor-toxin conjugate specifically inhibits KDR/Flk-1positive<br />

endothelial cell proliferation in vitro and angiogenesis in vivo, Cancer Res., 56, 1324, 1996.<br />

Ran, S. et al., Infarction of solid Hodgkin’s tumors in mice <strong>by</strong> antibody-directed targeting of tissue factor to<br />

tumor vasculature, Cancer Res., 58, 4646, 1998.<br />

Rettig, W. J. et al., Identification of endosialin, a cell surface glycoprotein of vascular endothelial cells in<br />

human cancer, Proc. Natl Acad. Sci. USA, 89, 10832, 1992.<br />

Ruoslahti, E., RGD and other recognition sequences for integrins, Annu. Rev., Cell Dev. Biol., 12, 697, 1996.<br />

Ruoslahti, E. et al., Alignment of biologically active domains in the fibronectin molecule, J. Biol. Chem., 256,<br />

7277, 1981.<br />

Saiki, I. et al., Inhibition of the metastasis of murine malignant melanoma <strong>by</strong> synthetic polymeric peptides<br />

containing core sequences of cell-adhesive molecules, Cancer Res., 49, 3815, 1989.<br />

Saiki, I. et al., Anti-metastatic and anti-invasive effects of polymeric Arg-Gly-Asp (RGD) peptide, poly<br />

(RGD), and its analogues, Jpn. J. Cancer Res., 81, 660, 1990.<br />

Sato, T.N. et al., Distinct roles of the receptor tyrosine kinases Tie-1 and Tie-2 in blood vessel formation,<br />

Nature, 376, 70, 1995.<br />

Scarborough, R. M. and Gretler, D. D., Platelet glycoprotein IIb-IIIa antagonists as prototypical integrin<br />

blockers: Novel parenteral and potential oral antithrombotic agents, J. Med. Chem., 43, 3453, 2000.<br />

Schiffelers, R. M. et al., Anti-tumor efficacy of tumor vasculature-targeted liposomal doxorubicin, J. Control.<br />

Release, 91, 115, 2003.<br />

Schlingemann, R. O. et al., Leukocyte antigen CD34 is expressed <strong>by</strong> a subset of cultured endothelial cells and<br />

on endothelial abluminal microprocesses in the tumor stroma, Lab. Investig., 62, 690, 1990.<br />

Seon, B. K. et al., Long-lasting complete inhibition of human solid tumors in SCID mice <strong>by</strong> targeting<br />

endothelial cells of tumor vasculature with antihuman endoglin immunotoxin, Clin. Cancer Res.,<br />

3,1031,1997.<br />

Shimaoka, M., Takagi, J., and Springer, T. A., Conformational regulation of integrin structure and function,<br />

Annu. Rev. Biophys. Biomol. Struct., 31, 485, 2002.<br />

Springer, T. A., Traffic signals for lymphocyte recirculation and leukocyte emigration: The multi-step paradigm,<br />

Cell, 76, 301, 1994.<br />

Suzuki, S. and Naitoh, Y., Amino acid sequence of a novel integrin 34 subunit and primary expression of the<br />

mRNA in epithelial cells, EMBO J., 9, 757, 1990.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


RGD-Modified Liposomes for Tumor Targeting 661<br />

Tabata, M. et al., Antiangiogenic radioimmunotherapy of human solid tumors in SCID mice using (125)Ilabeled<br />

anti-endoglin monoclonal antibodies, Int. J. Cancer, 82, 737, 1999.<br />

Tamura, R. N. et al., Epithelial integrin a4p6: Complete primary structure of a6 and variants forms of P4, J. Cell<br />

Biol., 111(1593), 604, 1990.<br />

Thurston, G. et al., Cationic liposomes target angiogenic endothelial cells in tumors and chronic inflammation<br />

in mice, J. Clin. Invest., 101,1401, 1998.<br />

Varner, J. A. and Cheresh, D. A., Tumor angiogenesis and the role of vascular cell integrin avb3, Important<br />

Adv. Oncol., 69, 1996.<br />

Vyas, S. P. and Khar, R. K., Drug Delivery to Tumor, in Targeted and Controlled Drug Delivery, Novel<br />

Carrier Systems, CBS Publishers, New Delhi, 2001.<br />

Xiong, X. B. et al., Enhanced intracellular uptake of sterically stabilized liposomal Doxorubicin in vitro<br />

resulting in improved antitumor activity in vivo, Pharm. Res., 22, 933, 2005.<br />

Yamamoto, S. et al., Antimetastatic effects of synthetic peptides containing the core sequence of the type III<br />

connecting segment domain (IIICS) of fibronectin, Anticancer Drugs, 5, 4424, 1994.<br />

Yusuf-Makagiansar, H. et al., Inhibition of LFA-1/ICAM-1 and VLA-4/VCAM-1 as a therapeutic approach to<br />

inflammation and autoimmune diseases, Med. Res. Rev., 22, 146, 2002.<br />

Zardi, L. et al., Elution of fibronectin proteolytic fragments from a hydroxyapatite chromatography column,<br />

Eur. J. BioChem., 146, 5711, 1985.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


33<br />

CONTENTS<br />

Folate Receptor-Targeted<br />

Liposomes for Cancer Therapy<br />

Xiaobin B. Zhao, Natarajan Muthusamy,<br />

John C. Byrd, and Robert J. Lee<br />

33.1 Introduction ....................................................................................................................... 663<br />

33.2 The Folate Receptor as a Cancer-Specific Cellular Marker............................................. 663<br />

33.3 Targeting FR for Cancer Therapy..................................................................................... 664<br />

33.4 Liposomes as Drug Carriers.............................................................................................. 665<br />

33.5 Formulation Strategies for FR-Targeted Liposomes ........................................................ 667<br />

33.6 In Vitro Drug Delivery Using FR-Targeted Liposomes .................................................. 668<br />

33.7 Pharmacokinetic and Biodistribution Properties of FR-Targeted Liposomes ................. 670<br />

33.8 In Vivo Therapeutic Efficacy of FR-Targeted Liposomes ............................................... 671<br />

33.9 Conclusions and Future Directions ................................................................................... 672<br />

Acknowledgments ........................................................................................................................ 673<br />

References..................................................................................................................................... 673<br />

33.1 INTRODUCTION<br />

Targeted drug delivery is a promising strategy to improve both the efficacy and safety of treatment.<br />

This is especially attractive for cancer therapy as most anti-cancer drugs have only marginal<br />

therapeutic index. Various cellular surface receptors and antigens have been targeted through<br />

different strategies, among which folate receptor (FR) has been one of the targets extensively<br />

investigated. FR is selectively amplified on human malignant cells and can take up folate and its<br />

analogs through a receptor-mediated endocytosis process. 1,2 FR-targeted liposomes can be utilized<br />

to deliver therapeutic agents to FR positive cells. 3 These liposomes have been evaluated for the<br />

targeted delivery of a wide variety of agents. This review will highlight recent developments on<br />

FR-targeted liposomes with emphasis on the application of this targeted nanoscale delivery system<br />

in cancer therapy. Potential applications of the FR-targeted liposomes in the clinics will also<br />

be discussed.<br />

33.2 THE FOLATE RECEPTOR AS A CANCER-SPECIFIC CELLULAR MARKER<br />

FR, also known as folate-binding proteins (FBP), is an N-glycosylated protein with high binding<br />

affinity to folate. FR has three isoforms, a, b, and g/g 0 . FR-a and FR-b are glycosyl-phosphatidylinositol<br />

(GPI)-anchored membrane bound proteins, 4,5 whereas FR-g/g 0 are constitutively<br />

secreted. 1,6,7 The g 0 isoform is a truncated form of FR-g. 8 These isoforms display different patterns<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

663


664<br />

of tissue specificity, and they also present differential ligand stereospecificities. 1 The two<br />

membrane receptor subtypes, a and b, share high amino acid sequence identity (70%) but are<br />

distinguishable <strong>by</strong> differential affinities for folic acid and stereoisoforms of reduced folates (affinities<br />

for folic acid: FR-a K dw0.1 nM, 9 FR- b K dw1 nM, 10 and FR- g K dw0.4 nM 6 ).<br />

Expression of FR is both tissue-specific and differentiation dependent. The expression of FR<br />

has been identified <strong>by</strong> immuno-histochemical staining, reverse transcription-polymerase chain<br />

reaction (RT–PCR), Western blotting, and 3 H-folic acid binding both in normal and malignant<br />

tissues. 11,12 Functional FR expression is absent in most normal tissues with the exception of the<br />

luminal surface of certain epithelial cells 2 where it has limited accessibility to the blood stream. On<br />

the other hand, FR-a is consistently expressed in several carcinomas, 12 especially in non-mucinous<br />

ovarian carcinomas, uterine carcinomas, testicular choriocarcinomas, ependymomas, and pleural<br />

mesotheliomas, and it is less frequently expressed in breast, colon, and renal cell carcinomas. 2,13<br />

Correlation of FR-a expression level and histologic grade has also been shown in ovarian and<br />

breast cancers, suggesting FR-a as a possible malignant cellular marker for cancers. 14 Methods for<br />

upregulating FR-a expression using anti-estrogens and gluococorticoid agonists have recently been<br />

published. 15,16 The effect of these agents is further enhanced <strong>by</strong> histone deacetylase (HDAC)<br />

inhibitors. 16 FR-b is a differentiation marker in the myelomonocytic lineage during neutrophil<br />

maturation, 17 and it is amplified in activated monocytes and macrophages. 18 However, FR-b in<br />

neutrophils is unable to bind folate as a result of aberrant post-translational modifications. 19 FR-b<br />

is also expressed in a functional form in chronic myelogenous leukemia (CML), in 70% of acute<br />

myelogenous leukemias (AML), 7,17,20,21 and in activated macrophages associated with rheumatoid<br />

arthritis. 22 FR-b expression is regulated <strong>by</strong> retinoid receptors and can be upregulated <strong>by</strong> all-trans<br />

retinoic acid (ATRA). 23 The effect of ATRA is further enhanced <strong>by</strong> HDAC inhibitors. The selective<br />

amplification of FR-a and FR-b expression in human cancers suggests possible roles for FRs<br />

as cancer specific cellular markers and their potential utility as targets for drug and gene delivery<br />

to these diseases. FR-g/g 0 are reported to be specific to hematopoietic tissues but are expressed<br />

at very low levels. 7 FR-g/g 0 are in secreted form, having limited application for targeted drug<br />

delivery but may serve as a serum marker to monitor certain hematological malignancies. 7 The<br />

additional possibility of selective FR upregulation in the target cells might further enhance the<br />

targeting strategy.<br />

33.3 TARGETING FR FOR CANCER THERAPY<br />

Nanotechnology for Cancer Therapy<br />

The selective amplification of FR expression in both human solid tumors and leukemia suggests its<br />

utility as a potentially valuable target for drug and gene delivery. Both monoclonal antibodies (MAbs)<br />

against FR and folate itself have been evaluated as moieties for targeting to the FR.<br />

Two MAbs, MOv18, and MOv19, have been produced against FR-a on ovarian cancers. Both<br />

aremurineMAbsoftheIgG1 class raised against a poorly differentiated ovarian carcinoma,<br />

recognizing two distinct epitopes. 24 Clinical studies on radioimmunoscintigraphy using<br />

131 I-MOv18 have been carried out in ovarian cancer patients and demonstrated targeting of<br />

ovarian cancer. 25 In addition, a particle-emitter 211 At conjugated to MOv18 has been found to<br />

prolong survival in a murine peritoneal tumor model developed <strong>by</strong> intraperitoneal inoculation of<br />

OVCAR-3 cells.<br />

Folate itself has, in fact, been the focus of recent development of FR targeting ligand. This is<br />

possibly because folate is a low molecular weight (MWZ441) ligand and has high affinity to FR,<br />

the conjugation chemistry of folate is defined and relatively easy to carry out, folate is a compound<br />

of unlimited availability, and folate lacks immunogenicity in contrast to murine MAb. The conjugation<br />

of folate via its g-carboxyl has been shown to preserve its binding affinity to the FRs. Recent<br />

studies also found that the glutamate residue of folic acid is not critical for FR recognition, and only<br />

pteroic acid is essential for FR binding. 61 Similar to the uptake process of the vitamin folate, FRs<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Folate Receptor-Targeted Liposomes for Cancer Therapy 665<br />

mediate the cellular internalization of folate conjugates via receptor-mediated endocytosis; 19,20<br />

therefore, this highly specific event can be utilized to promote uptake of folate-therapeutic agent<br />

conjugates in FR positive cancer cells.<br />

Early proof-of-concept in vivo studies were conducted using either radio-labeled or fluorochrome-labeled<br />

folic acid. The selectivity has also been documented in tumor-bearing mice using<br />

radiolabeled folate conjugates such as 67 Ga-deferoxamine-folate, 99m Tc-hydrazinonicotinic acid<br />

(HYNIC)-folate, and 111 In-diethylenetriamine pentaacetic acid (DTPA)-folate as potential folatebased<br />

radiopharmaceuticals. 26,27 Among these, 111 In-DTPA-folate has been further evaluated in<br />

human clinical trials. Concentration of radioactivity in abdominal masses was found in women<br />

suspected of ovarian cancer, and only the kidneys and livers in some patients displayed significant<br />

retention of 111 In-DTPA-folate. This reagent was developed <strong>by</strong> Endocyte Inc., an Indiana-based<br />

biopharmaceutical company, to be a folate-targeted radiopharmaceutical imaging agent<br />

(FolateScane) and is currently in phase II clinical trial designed to evaluate if FolateScane can<br />

be used to detect and make a disease or pathological assessment of ovarian cancer.<br />

A wide range of therapeutic agents have also been evaluated for enhanced cancer cell selective<br />

delivery. These include chemotherapeutic agents, 28 radiopharmaceuticals, 29 prodrug-converting<br />

enzymes, 30 protein toxins, 31 T-cell specific antibodies, 32 haptens, 30 gene transfer vectors, 33,34<br />

antisense oligodeoxyribonucleotides (ODNs), 35 polymeric drug carriers, 36 and liposomes<br />

19,37,38,44,49,51,56–60 carrying a variety of therapeutic agents. These conjugates potentially<br />

have broad applications in numerous imaging and therapeutic modalities.<br />

Based on the action site of therapeutic agents, application of FR-targeted delivery can be<br />

categorized into two types. For hydrophilic compounds with limited permeability through cell<br />

membrane, FR offers an efficient drug uptake pathway that can both specify and facilitate the<br />

accessibility of therapeutic active agents to cancer cells. Folate-conjugates enter FR positive<br />

cells through receptor-mediated endocytosis that is different from passive permeation process.<br />

Delivery to cancer cells through a process different from traditional drug transportation has the<br />

potential to make the drug more accessible to cancer cells. On the other hand, for drugs that are<br />

ready to take effect on the surface of cells such as immunomodulator and prodrug-activating<br />

enzymes, FR may also act as a concentrator for these types of drugs. FRs are generally over<br />

expressed on the cellular surface of various cancer types, and upon binding of FR drug conjugates,<br />

they may enrich these therapeutic agents on the cell surface. Therefore, application of FR-targeted<br />

delivery can be exploited for variety of compounds (Figure 33.1).<br />

Taken together, FR targeting has been investigated as a valuable approach to enhance the<br />

delivery of variety of agents. Therefore, the interest in exploiting this receptor for targeting application<br />

is still increasing as demonstrated <strong>by</strong> the active researches conducted in both academia and<br />

industry. Among these, FR-targeted liposome, first designed and investigated <strong>by</strong> Lee and Low, 37,38<br />

has been attractive because of the advantages it possesses as a useful delivery system both for<br />

hydrophilic and hydrophobic agents.<br />

33.4 LIPOSOMES AS DRUG CARRIERS<br />

Liposomes are spherical, self-closed structures composed of lipophilic bilayers with entrapped<br />

hydrophilic core. The development of small size, stable, long-circulating, targeted liposomes has<br />

led to an exciting era in the pharmaceutical application of nanotechnology. Giving the unique<br />

structure of liposomes compared to other drug delivery systems, both lipophilic and lipophobic<br />

therapeutic agents can be delivered <strong>by</strong> liposomes. The biocompatibility of liposomes, if used with<br />

proper lipid composition, also makes liposomes attractive as safe and effective drug carriers. A<br />

variety of therapeutic agents have been incorporated into liposomes. Several have reached clinical<br />

use. These include liposomal doxorubicin (Doxile), daunorubicin (Daunoxomee), amphotericin B<br />

(Amphotece, Ambisomee, Abelcete), cytarabine (Depocytee), and verteporfin (Visudynee).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


666<br />

OH<br />

N<br />

H N<br />

2N N<br />

N<br />

Folic acid<br />

NH<br />

α<br />

O<br />

COOH<br />

NH<br />

COOH γ<br />

FIGURE 33.1 Folate-conjugates for FR-targeted delivery.<br />

Numerous liposomal formulations are in clinical trial, including vincristine, all-trans retinoid acid,<br />

topotecan, and cationic liposome-based therapeutic gene transfer vectors. Many more are in preclinical<br />

evaluation. Besides potential use in systemic gene delivery, cationic liposomes are routinely<br />

used as transfection reagents for plasmid DNA and oligonucleotides in the laboratory. The use of<br />

liposomes can be deliberately engineered to possess unique properties such as long systemic<br />

circulation time, pH sensitivity, temperature sensitivity, and target cell specificity. These are<br />

achieved <strong>by</strong> selecting the appropriate lipid composition and surface modification of the liposomes,<br />

expanding the application of liposomes for drug delivery.<br />

Liposomal delivery of anti-cancer drugs has been shown to greatly extend their systemic<br />

circulation time, reduce toxicity <strong>by</strong> lowering plasma free drug concentration, and facilitate preferential<br />

localization of drugs in solid tumors based on increased endothelial permeability and reduced<br />

lymphatic drainage or enhanced permeability and retention (EPR) effect. 39 Clearance of drugs in a<br />

liposomal formulation is mediated <strong>by</strong> phagocytic cells of the reticuloendothelial system (RES),<br />

primarily located in the liver and spleen. Drugs can be loaded during liposome preparation through<br />

methods such as passive entrapment. Alternatively, they can also be loaded after liposome preparation<br />

through a process called remote loading, utilizing the ion gradient across the membrane as<br />

in the case of doxorubicin. 31 Compared to free drugs, liposomal drugs are not subjected to rapid<br />

renal clearance. Instead, they are cleared <strong>by</strong> RES, exhibiting prolonged systemic circulation time. 40<br />

RES clearance of liposomes can be reduced <strong>by</strong> surface coating with polyethyleneglycol (PEG).<br />

Passive targeting effect has also been shown in liposomal drug delivery to solid tumors where<br />

liposomes preferentially localize in these tumor tissues as a result of the EPR effect. 41<br />

Liposomes can be targeted to specific cell populations via incorporation of a desired targeting<br />

moiety such as a chemical conjugated ligand, antibody, or antibody fragment directed against the<br />

surface molecules. Targeted delivery of these liposomes can greatly improve their selectivity of<br />

cancer cells and facilitate their cellular uptake and intracellular drug release. Examples of such<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

E<br />

Nanotechnology for Cancer Therapy<br />

Radiopharmaceuticals<br />

Chemotherapeutics<br />

DNA and antisense ODN<br />

Prodrug converting enzymes<br />

Antibodies<br />

BNCT agents<br />

Polymeric drug carriers<br />

Liposomes


Folate Receptor-Targeted Liposomes for Cancer Therapy 667<br />

targeted liposomes are folate-conjugated liposomes targeting to acute myelogenous leukemia,<br />

CD19-targeted immunoliposomes for non-Hodgkin’s lymphoma (NHL) therapy, 42 and<br />

anti-HER2 immunoliposomal doxorubicin targeting to HER2-overexpressing breast cancer<br />

cells. 43 A variety of techniques have been described to incorporate targeting moieties to<br />

liposomes. First, ligands such as folate can beconjugatedtoalipidcomponent suchas<br />

poly(ethylene glycol)-conjugated distearoyl phosphatidylethanolamine (PEG-DSPE) and PEG-<br />

Chol. This amphophilic lipid component can be constructed into the lipid bilayer during the<br />

formation of membrane leaving the ligand on the surface of the particles. 38,59 In the second<br />

method, a functionalized lipid anchor is included first to the outside of the liposomes after the<br />

formation of the bilayer membrane. Therefore, the exposed crosslinker can react with activated<br />

targeting moieties such as thiolated proteins (antibody, scFv, transferrin, etc.). 42,43 A useful thiolation<br />

reagent is 2-iminothiolane that is also known as Traut’s reagent. There are two types of<br />

immunoliposomes based on the antibody coupling strategy. Type one involves direct attachment<br />

of the antibodies to the lipids, and type two involves linking the antibodies to the terminal ends of<br />

reactive PEG derivates. Most of these coupling techniques lead to a random antibody orientation at<br />

the liposome surface. A significant contribution to the development of the type two immunoliposome,<br />

recently adopted for formulation of HER2-targeted immunoliposomes, is an antibody<br />

coupling method called post-insertion method. 43 This method involves formation of micelles of<br />

lipid-derivatized antibodies first, followed <strong>by</strong> transferring the antibody-coupled PEG-lipid micelles<br />

to the preformed liposomes, converting the conventional liposomes to immunoliposome through a<br />

simple one-step incubation method. Maleimide-terminated PEG–DSPE is a coupling lipid that has<br />

been validated to successfully convert commercially available liposomal doxorubicin (Doxil/<br />

Caelyx) to targeted sterically stabilized liposomal doxorubicin. The last post-insertion method<br />

seems highly promising for future clinical development.<br />

33.5 FORMULATION STRATEGIES FOR FR-TARGETED LIPOSOMES<br />

FR-targeted liposomes can be prepared <strong>by</strong> including a small fraction (e.g., 0.1 mole%) of a lipophilic<br />

folate derivative into the lipid composition. 38 Both folate-PEG (M.W. 3350)-distearoyl<br />

phosphatidylethanolamine (folate-PEG–DSPE) 38 and folate-PEG-cholesterol (folate-PEG-chol)<br />

have been synthesized and shown to be effective in targeting liposomes to the FR. The synthesis<br />

of folate-PEG–DSPE and folate-PEG-Chol can be realized through two stages. First, an intermediate<br />

folate-PEG-amine is synthesized <strong>by</strong> reacting NHS-activated folate and reacted with PEG-bis<br />

amine and further purifized using a Sepharose column. This folate-PEG-NH 2 is then ready to react<br />

with N-succinyl-DSPE to form folate-PEG–DSPE 38 or with cholestreryl chloroformate to form<br />

folate-PEG-chol. 59 Another synthetic method has also been described <strong>by</strong> Gabizon that is to couple<br />

FA to readily attainable H2N-PEG–DSPE with the mediation of carbodiimide. 44 These two synthesized<br />

lipophilic folate-derivatives have both been shown to be good components to be<br />

incorporated to liosomes. Compared to the synthesis of folate-PEG–DSPE, the synthesis of<br />

folate-PEG-chol is relatively simple and is less costly. With regard to stability, cholesterol also<br />

provides better chemical stability, and it is also known to be able to increase the structural integrity<br />

of lipid membranes through tight packing. The molecule of cholesterol is neutrally charged,<br />

different from the negatively charged DSPE molecule, and can possibly provide better targeting<br />

effect when used as a lipophilic anchor to incorporate the folate molecule. For these reasons, this lab<br />

has been routinely using folate-PEG-chol for the formulation of FR-targeted liposomes.<br />

Preparation of folate receptor (FR) targeted liposomes is relatively straightforward since the<br />

targeting lipid component can simply be included into the lipid composition during the formation of<br />

liposomes. A small percentage (0.1–0.5%, molar ratio) of folate-PEG–DSPE or folate-PEG-chol in<br />

the lipid can yield targeting effect to FR. Relatively long linker between folate and the lipid anchor<br />

(e.g., PEG 3350) was found to be necessary for FR binding of the liposomes. 37 FR targeting was<br />

compatible with incorporation of 3 mole% of PEG–DSPE that is required for prolonging the in vivo<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


668<br />

Nanotechnology for Cancer Therapy<br />

circulation time of the liposomes. Lab scale production of these targeted liposomes can be<br />

performed <strong>by</strong> thin-layer hydration followed <strong>by</strong> polycarbonate membrane extrusion. For scaling<br />

up, alternative methods such as liquid phase emulsification, high pressure homogenization, and<br />

tangential flow filtration can be used. Recently, this lab has successfully validated a processing<br />

method combining homogenization, ultrafiltation, and remote loading for FR-targeted liposomal<br />

doxorubicin at clinical relevant scale quantity. Using this method, three consecutive batches of<br />

liposomes with mean particle size around 100 nm curve produced. Another alternative method for<br />

folate-PEG-Chol incorporation has not yet been investigated, using post-insertion method similar to<br />

HER2-targeted immunoliposome production <strong>by</strong> incubating micelles of folate-PEG-chol/PEG-chol<br />

to pre-formed liposomal drugs. The latter method avoided the ligand stability concern during the<br />

liposome production and has the potential to be used for future Current Good Manufacturing<br />

Practices (cGMP) grade production of FR-targeted liposomes. Both hydrophilic drugs and hydrophobic<br />

drugs can be potentially loaded to the targeted particles through remote loading<br />

(doxorubicin, daunorubicin, vincristine, and topetecan) or passive entrapment (paclitaxel and<br />

ATRA); therefore, this FR-targeted liposomal drug delivery represents a valuable approach for<br />

FR-positive cancer therapy (Figure 33.2).<br />

FR-targeted cationic liposome formulations have also been studied for gene delivery. Several<br />

cationic liposome formulations that incorporate a lipopholic folate derivative as a targeting entity<br />

have been reported. FR-targeted lipoplexes constructed using a cationic lipid RPR209120 was<br />

shown to efficiently transfer gene. 45 Another study using 2% DPPE–PEG 3340-folate and a cationic<br />

dithiol-detergent (dimerized tetradecyl-ornithinly-cysteine) demonstrated efficient FR-dependent<br />

cellular uptake and transfection. Therapeutic gene p53 was also delivered <strong>by</strong> Xu et al., using<br />

cationic liposomes conjugated to folate, and this was shown to enhance the anti-tumor efficacy<br />

of conventional chemotherapy. 46 FR-targeted cationic liposomes can further form complex with<br />

DNA-condensing polymer to form FR-targeted lipopolyplexes, and this was demonstrated <strong>by</strong><br />

Reddy et al. 54 to have superior transfection activity in FRCM109 murine lung carcinoma cells as<br />

well as in L1210A murine lymphocytic leukemia cells. Lee et al. developed FR-targeted type II<br />

lipopolyplexes, and these FR-targeted vectors exhibited improved gene transfer activity even in the<br />

presence of serum. 47,48 Leamon et al. also investigated the use of FR-targeted liposome constructed<br />

with lipid-PEG-pteroate for effective delivery of antisense oligonucleotides into FR-bearing cells<br />

in vitro. 61 FR-targeted liposomal drug and gene delivery has also been the subject of another review<br />

article. 35<br />

33.6 IN VITRO DRUG DELIVERY USING FR-TARGETED LIPOSOMES<br />

Cellular uptake of FR-targeted liposomes has been characterized using KB cells, a FR-a (C)<br />

human oral carcinoma cell line. 37 Binding of the liposomes has a relatively slow kinetics, occurring<br />

over several hours, presumably a result of their size. The specific binding of these liposomes to KB<br />

cells was saturable and could be blocked <strong>by</strong> excess free folate in the binding media. 37 Interestingly,<br />

the apparent affinity of the folate-conjugated liposomes to KB cells was much higher than that of<br />

free folate. 37 This is due to multivalent binding of the folate ligand on the liposomes with the FRs<br />

on the cellular surface. 37 The folate-liposomes are internalized into a low pH compartment via<br />

receptor-mediated endocytosis, and the encapsulated drug molecules are released into the cytoplasm<br />

to induce cytotoxicity 47 as illustrated in Figure 33.3.<br />

Drug delivery properties of FR-targeted liposomes have been studied in vitro using liposomes<br />

loaded with chemotherapeutic agents such as doxorubicin, 38 daunorubicin, 49 and cisplatin. 44 Lee<br />

et al. first reported the in vitro effect of doxorubin and showed these targeted liposomes showed<br />

w86-fold greater cytotoxicity in KB cells compared to non-targeted control liposomes. 38 The<br />

enhancement in cytotoxicity was correlated with the increase in doxorubicin uptake and could<br />

be blocked <strong>by</strong> excess free folate. 38 Furthermore, a study <strong>by</strong> Goren et al. showed that the delivery of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Folate Receptor-Targeted Liposomes for Cancer Therapy 669<br />

H 2 N<br />

N<br />

H 2 N<br />

OH<br />

N<br />

N<br />

N<br />

N<br />

OH<br />

N<br />

NH<br />

N<br />

N<br />

NH<br />

Folate<br />

Folate receptor<br />

HOOC<br />

NH<br />

O<br />

O<br />

Folate-PEG-cholesterol<br />

COOH PEG<br />

O<br />

NH NH<br />

NH<br />

O<br />

O<br />

NH<br />

O<br />

Folate-PEG-DSPE<br />

PEG 2000<br />

PEG O<br />

NH<br />

NH<br />

O<br />

O P O<br />

Hydrophobic reagents<br />

Hydrophilic reagents<br />

FR-targeted liposome<br />

Cancer cell membrane<br />

FIGURE 33.2 FR-targeted liposome. The ligand folate is conjugated to a lipophilic anchor (PEG3350-DSPE or<br />

PEG3350-chol) and incorporated to the lipid bilayer. PEG2000–DSPE is also included in the construction of the<br />

liposome membrane, prolonging the systemic circulation of these particles <strong>by</strong> avoiding the RES uptake.<br />

Hydrophilic anti-cancer reagents can be loaded to the aqueous core through active loading or passive entrapment,<br />

and hydrophobic reagents can be loaded to the lipid bilayer through formation of electrostatic complex<br />

or passive entrapment. The structures of two synthetic targeting lipids (F-PEG–DSPE and F-PEG-Chol) were<br />

also shown.<br />

doxorubicin via FR-targeted liposomes <strong>by</strong>passed Pgp-dependent drug efflux in drug resistant<br />

FR-a(C) M109 murine lung carcinoma cells in vitro and in an adoptive assay in mice. 50 These<br />

data suggest that this delivery modality might overcome drug resistance in these tumor cells. 50<br />

More recently, folate-liposomal doxorubicin was studied in FR-b(C) KG1 human AML cells and<br />

similarly showed enhanced cytotoxicity relative to non-targeted control liposomes, 19 and the effect<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O-<br />

O<br />

O<br />

O<br />

O<br />

O


670<br />

FIGURE 33.3 FR-mediated endocytosis of folate-liposomes.<br />

was enhanced <strong>by</strong> FR-b upregulation using ATRA. 19 Therefore, both FR-a and FR-b are potential<br />

targets for folate-conjugated liposomes.<br />

In addition to chemotherapeutics, a variety of agents have been loaded into FR-targeted<br />

liposomes and evaluated in vitro. These include fluorescent 37 probes, BNCT agents, 51 photosensitizers,<br />

52 antisene oligodeoxyribonucleotides against the EGF receptor, 53 and plasmid DNA<br />

carrying luciferase reporter gene. 47,54 Similar formulations of FR-targeted lipid nanoparticles<br />

containing paclitaxel have also been recently reported. 55 In each of these studies, the FR-targeted<br />

formulationwasshowntobemuchmoreefficient in cellular uptake and biological activity<br />

compared to the non-targeted control formulation.<br />

33.7 PHARMACOKINETIC AND BIODISTRIBUTION PROPERTIES<br />

OF FR-TARGETED LIPOSOMES<br />

Nanotechnology for Cancer Therapy<br />

The in vivo clearance rate for folate-conjugated liposomal doxorubicin was faster than non-targeted<br />

control liposomes in mice engrafted with FR-overexpressing M109 or human KB carcinoma or<br />

mouse J6453 lymphoma. 56 These results also suggested a possible role of FR in the differential<br />

rates of clearance. 56 Plasma clearance kinetics of FR-targeted liposomes radiolabled with 67 Ga<br />

were studied in rats. These liposomes showed enhanced plasma clearance rates and higher uptakes<br />

in the liver and spleen compared to non-targeted control liposomes. 56 This effect was even more<br />

pronounced in rats that were placed on a folate-free diet, suggesting that FR expression in the RES<br />

organs might be responsible for the increase in the rate of liposome clearance. 56 One possible explanation<br />

is that FR-targeted liposomes, <strong>by</strong> virtue of their multivalency, can interact with the FR-b that<br />

is present at a relatively low level in the phagocytic cells.<br />

In the biodistribution studies in murine tumor models, FR-targeted liposomes did not show<br />

enhanced accumulation in tumors relative to non-targeted liposomes. 59 This might be due to the<br />

passive accumulation of the liposomes via the EPR effect was the predominant factor in tumor<br />

localization of both targeted and non-targeted liposomes because endothelial extravasation was<br />

the rate-limiting step in tumor localization. In addition, FR-mediated enhancement in uptake of<br />

the targeted liposomes might be offset <strong>by</strong> the more rapid plasma clearance of these liposomes as<br />

discussed above. 60 Moreover, in solid tumors, FRs on tumor cell surface are not directly exposed<br />

to the circulation and not fully accessible <strong>by</strong> circulating FR-targeted liposomes. Therefore, extravasation,<br />

instead of specific binding, is the rate-limiting factor of liposome distribution in<br />

solid tumors.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Folate Receptor-Targeted Liposomes for Cancer Therapy 671<br />

An important observation in vivo is that macrophages also have high uptake of FRtargeted<br />

liposomes. In a study done in an ovarian carcinoma ascites mouse model, results<br />

showed that macrophages took up FR-targeted liposomes efficiently in the ascites fluid even in<br />

the presence of FR positive ascetic tumor cells, 57 although other reports did show increased<br />

FR-targeted liposome uptake in FR(C) ascitic tumor cells. 56 The relative uptake of folateliposomes<br />

<strong>by</strong> these two cell populations was likely determined <strong>by</strong> the relative FR expression<br />

levels. 56<br />

33.8 IN VIVO THERAPEUTIC EFFICACY OF FR-TARGETED LIPOSOMES<br />

The therapeutic efficacy of folate-liposomal doxorubicin has been evaluated in several murine<br />

tumor models. In a KB cell xenograft model of athymic mice on a folate-free diet, folate-liposomal<br />

doxorubicin given i.p. at 10 mg/kg!6 injections showed significantly greater tumor inhibition<br />

compared to non-targeted liposomes. 58 However, in another study using a M109R-HiFR syngraft<br />

model in BALB/c mice on a regular folate-containing diet, folate-liposomal doxorubicin given<br />

intravenously at 8 mg/kg!4 injections showed a similar therapeutic efficacy to non-targeted<br />

control liposomes. 3 The differences in these studies might be partly due to the different routes of<br />

administration, animal diet, and dosages applied. 3 A challenge in demonstrating a therapeutic<br />

benefit of folate-liposomal doxorubicin is the excellent anti-tumor activity of PEGylated liposomal<br />

doxorubicin used as a control that has optimal pharmacokinetic properties and EPR effect and are<br />

much more effective than the free drug. 3 Further studies in additional models are likely needed to<br />

confirm possible therapeutic benefits of FR-targeted liposomal doxorubicin.<br />

A possible mechanism for increased therapeutic efficacy for the folate-liposomal doxorubicin in<br />

solid tumor is altered intratumoral drug distribution. These liposomes might be more efficiently<br />

internalized <strong>by</strong> FR(C) tumor cells and/or tumor infiltrating macrophages that are FR(C). 57 This, in<br />

turn, results in an increase in drug release within the tumor. In contrast, non-targeted liposomal<br />

doxorubicin may remain extracellular and be distributed in the interstitial space and releases drug<br />

more slowly. This may result in a bioavailable (free) drug concentration that falls below the<br />

minimum effective concentration (MEC) for inducing tumor cell apoptosis. Therefore, cellular<br />

internalization and associated drug release may be a critical mechanism of liposomal drug action<br />

in solid tumors and a rate-limiting factor in overall drug delivery pathway. The relative importance<br />

of FR expression on tumor cells and tumor infiltrating macrophages and their role in uptake of<br />

folate-coated liposomes warrant further investigation.<br />

Leukemia is a better disease target as compared to solid tumors for this targeted delivery<br />

strategy. This potentially is because malignant cells are circulating in leukemia patients and are<br />

fully accessible <strong>by</strong> targeted liposomes. In contrast, these liposomes need to overcome several<br />

barriers in solid tumors before reaching cancer cells. Several studies have already reported the<br />

anti-tumor efficacy of FR-targeted liposomal anthracyclines in leukemia. In both L1210JF murine<br />

lymphocytic leukemia ascites models and a KG-1 human acute myelogenous leukemia xenograft<br />

model, FR-targeted liposomal doxorubicin or daunorubicin, when administered i.p, 19 showed<br />

improvement in survival of treated mice compared to non-targeted control liposomes or the free<br />

drug in the L1210JF model. 19 In both mice models, folate-liposomal doxorubicin given i.p. was<br />

also more effective in prolonging animal survival. 19 The relatively greater accessibility of<br />

leukemia cells in ascites tumors might have contributed to the superior efficacy of the<br />

targeted liposomes.<br />

Interestingly, pretreatment of mice with ATRA, an agent known to upregulate FR-b expression<br />

in KG-1 cells in vitro, 19 before the injection of liposomes was also shown to enhance the<br />

efficacy. An increase in long-term survival in these mice as a result of FR-targeted liposomal<br />

doxorubicin was shown from 12.5% to 60%. Therefore, administrating ATRA to upregulate<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


672<br />

FR-b expression seems to be a potentially promising therapeutic strategy to be combined with the<br />

FR-targeted therapy.<br />

33.9 CONCLUSIONS AND FUTURE DIRECTIONS<br />

Nanotechnology for Cancer Therapy<br />

Targeted drug delivery is a promising approach to selectively deliver neoplastic agents to cancer<br />

cells and to improve the therapeutic index of chemotherapy. FR outstands among various tumor<br />

markers because of the specificity and relatively high level of endocytosis, representing a very<br />

attractive target for developing targeted delivery for future cancer therapy. FR-targeted liposomal<br />

delivery is a promising strategy for treatment of FR(C) tumors and leukemias. Although a variety<br />

of agents have been delivered to FR(C) tumor cell in vitro using these liposomes, relatively few<br />

reports address the in vivo properties of these liposomes. However, at least some studies indicated a<br />

therapeutic advantage of folate-liposomal doxorubicin over the non-targeted control. 19,49,58<br />

Leukemia, that has greater accessibility from the systemic circulation, may present an even<br />

better therapeutic target for these liposomes. The concept of sensitizing AML cells with<br />

FR-b-upregulating agents such as ATRA may bring further rationale to take FR-targeted liposomal<br />

drug delivery to clinical trials.<br />

Several critical factors to be considered for designing in vivo application of FR-targeted liposomes<br />

for cancer therapy include FR expression level on tumor cells relative to normal cells, the<br />

accessibility of liposomes to these receptors, the intra-tumor diffusion rate of vectors, and the<br />

presence of endogenous folate in systemic circulation. The first obstacle of insufficient levels of<br />

FR expression on tumor cells in patients could be possibly overcome <strong>by</strong> inducing FR expression<br />

upregulation <strong>by</strong> co-administration of anti-estrogen or retinoid receptor ligands. To address the<br />

second concern of accessibility, optimization of formulation parameter is necessary to reduce the<br />

particle size, reduce the plasma protein binding, and avoid the RES and other normal tissue uptake.<br />

For solid tumor delivery, priming the vasculature in the tumor to increase the distribution of<br />

targeted particles is a possible approach. The concern that the presence of endogenous folate in<br />

systemic circulation can block FR binding is not a critical one. It is believed that this should not<br />

constitute a barrier as the endogenous form of folate has lower affinity for the FR compared to folic<br />

acid and should not significantly impede the binding of multivalent folate-coated liposomes.<br />

Clinical success in targeted drug delivery area has so far been limited <strong>by</strong> the lack of a suitable<br />

cellular target and/or a high degree of difficulty in the production of clinical quantities of targeted<br />

drug carriers. FR-targeted liposomes represent an attractive candidate for clinical development with<br />

potential application in the treatment of AML and other FR-over expressing solid tumors. Areas of<br />

investigation for future studies include the possible role of receptor upregulation in promoting<br />

targeted therapy, both for FR-a and FR-b positive tumors; and the cellular population responsible<br />

for enhancement in therapeutic response in solid tumors, i.e., relative roles of tumor infiltrating<br />

macrophages and the tumor cells themselves. 57 Because activated macrophages have increased FRb<br />

expression, 22 it is possible that folate-conjugates can effectively target tumor infiltrating macrophages<br />

in non-FR expressing tumors, extending the FR-targeting strategy to receptor<br />

negative tumors.<br />

The area of FR targeting remains an area of great excitement, suggested <strong>by</strong> large number of<br />

recent publications and reports from industrial development. So far, no clinical study targeting FR<br />

for therapeutic purposes has been carried out. Given the advantages of the FR-targeting strategy,<br />

further preclinical studies are urgently needed to define mechanisms of in vivo targeting and the<br />

potential for clinical efficacy. FR-targeted liposomal doxorubicin represents an compelling candidate<br />

for clinical development with potential application in the treatment of AML. Furthermore,<br />

future application to other tumors with high frequency of FR overexpression such as ovarian and<br />

lung carcinomas warrants investigation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Folate Receptor-Targeted Liposomes for Cancer Therapy 673<br />

ACKNOWLEDGMENTS<br />

This work was supported, in part, <strong>by</strong> NCI grant ROI CA095673 to RJL, P01CA95426 to JCB, NSF<br />

grants EEC-0425626 to RJL, and Leukemia and Lymphoma Society grant 6113-02 to RJL.<br />

REFERENCES<br />

1. Elnakat, H. and Ratnam, M., Distribution, functionality and gene regulation of folate receptor<br />

isoforms: Implications in targeted therapy, Advanced Drug Delivery Reviews, 56(8), 1067–1084,<br />

2004.<br />

2. Weitman, S. D. et al., Distribution of the folate receptor GP38 in normal and malignant cell lines and<br />

tissues, Cancer Research, 52(12), 3396–3401, 1992.<br />

3. Gabizon, A. et al., Tumor cell targeting of liposome-entrapped drugs with phospholipid-anchored<br />

folic acid-PEG conjugates, Advanced Drug Delivery Reviews, 56(8), 1177–1192, 2004.<br />

4. Lacey, S. W. et al., Complementary DNA for the folate-binding protein correctly predicts anchoring<br />

to the membrane <strong>by</strong> glycosyl-phosphatidylinositol, The Journal of Clinical Investigation, 84(2),<br />

715–720, 1989.<br />

5. Ratnam, M. et al., Homologous membrane folate-binding proteins in human placenta: Cloning and<br />

sequence of a cDNA, Biochemistry, 28(20), 8249–8254, 1989.<br />

6. Shen, F. et al., Folate receptor type gamma is primarily a secretary protein due to lack of an efficient<br />

signal for glycosyl-phosphatidylinositol modification: Protein characterization and cell type specificity,<br />

Biochemistry, 34(16), 5660–5665, 1995.<br />

7. Shen, F. et al., Identification of a novel folate receptor, a truncated receptor, and receptor type beta in<br />

hematopoietic cells: cDNA cloning, expression, immunoreactivity, and tissue specificity, Biochemistry,<br />

33(5), 1209–1215, 1994.<br />

8. Wang, H., Ross, J. F., and Ratnam, M., Structure and regulation of a polymorphic gene encoding<br />

folate receptor type gamma/gamma 0 , Nucleic Acids Research, 26(9), 2132–2142, 1998.<br />

9. Kamen, B. A. and Caston, J. D., Properties of a folate-binding protein (FBP) isolated from porcine<br />

kidney, Biochemical Pharmacology, 35(14), 2323–2329, 1986.<br />

10. da Costa, M. and Rothenberg, S. P., Purification and characterization of folate-binding proteins from<br />

rat placenta, Biochimica Et Biophysica Acta, 1292(1), 23–30, 1996.<br />

11. Ross, J. F., Chaudhuri, P. K., and Ratnam, M., Differential regulation of folate receptor isoforms in<br />

normal and malignant tissues in vivo and in established cell lines, Physiologic and clinical implications.<br />

Cancer, 73(9), 2432–2443, 1994.<br />

12. Weitman, S. D. et al., Cellular localization of the folate receptor: Potential role in drug toxicity and<br />

folate homeostasis, Cancer Research, 52(23), 6708–6711, 1992.<br />

13. Garin-Chesa, P. et al., Trophoblast and ovarian cancer antigen LK26. Sensitivity and specificity in<br />

immunopathology and molecular identification as a folate-binding protein, American Journal of<br />

Pathology, 142(2), 557–567, 1993.<br />

14. Toffoli, G. et al., Overexpression of folate-binding protein in ovarian cancers, International Journal of<br />

Cancer, 74(2), 193–198, 1997.<br />

15. Kelley, K. M. M., Rowan, B. G., and Ratnam, M., Modulation of the folate receptor alpha gene <strong>by</strong> the<br />

estrogen receptor: Mechanism and implications in tumor targeting, Cancer Research, 63(11),<br />

2820–2828, 2003.<br />

16. Tran, T. et al., Enhancement of folate receptor alpha expression in tumor cells through the glucocorticoid<br />

receptor: A promising means to improved tumor detection and targeting, Cancer Research,<br />

65(10), 4431–4441, 2005.<br />

17. Ross, J. F. et al., Folate receptor type beta is a neutrophilic lineage marker and is differentially<br />

expressed in myeloid leukemia, Cancer, 85(2), 348–357, 1999.<br />

18. Nakamura, M., Nitration and chlorination of folic acid <strong>by</strong> peroxynitrite and hypochlorous acid, and the<br />

selective binding of 10-nitro-folate to folate receptor beta, Biochemical and Biophysical Research<br />

Communications, 297(5), 1238–1244, 2002.<br />

19. Pan, X. Q. et al., Strategy for the treatment of acute myelogenous leukemia based on folate receptor<br />

beta-targeted liposomal doxorubicin combined with receptor induction using all-trans retinoic acid,<br />

Blood, 100(2), 594–602, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


674<br />

Nanotechnology for Cancer Therapy<br />

20. Sadasivan, E. et al., Purification, properties, and immunological characterization of folate-binding<br />

proteins from human leukemia cells, Biochimica Et Biophysica Acta, 925(1), 36–47, 1987.<br />

21. Sadasivan, E. et al., Characterization of multiple forms of folate-binding protein from human<br />

leukemia cells, Biochimica Et Biophysica Acta, 882(3), 311–321, 1986.<br />

22. Paulos, C. M. et al., Folate receptor-mediated targeting of therapeutic and imaging agents to activated<br />

macrophages in rheumatoid arthritis, Advanced Drug Delivery Reviews, 56(8), 1205–1217, 2004.<br />

23. Wang, H. et al., Differentiation-independent retinoid induction of folate receptor type beta, a potential<br />

tumor target in myeloid leukemia, Blood, 96(10), 3529–3536, 2000.<br />

24. Miotti, S. et al., Characterization of human ovarian carcinoma-associated antigens defined <strong>by</strong> novel<br />

monoclonal antibodies with tumor-restricted specificity, International Journal of Cancer, 39(3),<br />

297–303, 1987.<br />

25. Kalofonos, H. P., Karamouzis, M. V., and Epenetos, A. A., Radioimmunoscintigraphy in patients with<br />

ovarian cancer, Acta Oncology, 40(5), 549–557, 2001.<br />

26. Mathias, C. J., et al., Tumor-selective radiopharmaceutical targeting via receptor-mediated endocytosis<br />

of gallium-67-deferoxamine-folate, Journal of Nuclear Medicine, 37(6), 1003–1008, 1996.<br />

27. Mathias, C. J., et al., Indium-111-DTPA-folate as a potential folate-receptor-targeted radiopharmaceutical,<br />

Journal of Nuclear Medicine, 39(9), 1579–1585, 1998.<br />

28. Leamon, C. P. and Reddy, J. A., Folate-targeted chemotherapy, Advanced Drug Delivery Reviews,<br />

56(8), 1127–1141, 2004.<br />

29. Ke, C. Y., Mathias, C. J., and Green, M. A., Folate-receptor-targeted radionuclide imaging agents,<br />

Advanced Drug Delivery Reviews, 56(8), 1143–1160, 2004.<br />

30. Lu, J. Y. et al., Folate-targeted enzyme prodrug cancer therapy utilizing penicillin-V amidase and a<br />

doxorubicin prodrug, Journal of Drug Targeting, 7(1), 43–53, 1999.<br />

31. Leamon, C. P. and Low, P. S., Selective targeting of malignant cells with cytotoxin-folate conjugates,<br />

Journal of Drug Targeting, 2(2), 101–112, 1994.<br />

32. Roy, E. J. et al., Folate-mediated targeting of T cells to tumors, Advanced Drug Delivery Reviews,<br />

56(8), 1219–1231, 2004.<br />

33. Mislick, K. A. et al., Transfection of folate-polylysine DNA complexes: Evidence for lysosomal<br />

delivery, Bioconjugate Chemistry, 6(5), 512–515, 1995.<br />

34. Gottschalk, S. et al., Folate receptor mediated DNA delivery into tumor cells: Potosomal disruption<br />

results in enhanced gene expression, Gene Therapy, 1(3), 185–191, 1994.<br />

35. Zhao, X. B. and Lee, R. J., Tumor-selective targeted delivery of genes and antisense oligodeoxyribonucleotides<br />

via the folate receptor, Advanced Drug Delivery Reviews, 56(8), 1193–1204,<br />

2004.<br />

36. Shukla, S. et al., Synthesis and biological evaluation of folate receptor-targeted boronated PAMAM<br />

dendrimers as potential agents for neutron capture therapy, Bioconjugate Chemistry, 14(1), 158–167,<br />

2003.<br />

37. Lee, R. J. and Low, P. S., Delivery of liposomes into cultured KB cells via folate receptor-mediated<br />

endocytosis, Journal of Biological Chemistry, 269(5), 3198–3204, 1994.<br />

38. Lee, R. J. and Low, P. S., Folate-mediated tumor cell targeting of liposome-entrapped doxorubicin<br />

in vitro, Biochimica Et Biophysica Acta, 1233(2), 134–144, 1995.<br />

39. Gabizon, A. and Papahadjopoulos, D., Liposome formulations with prolonged circulation time in<br />

blood and enhanced uptake <strong>by</strong> tumors, Proceedings of the National Academy of Sciences of the United<br />

States of America, 85(18), 6949–6953, 1988.<br />

40. Abraham, S. A. et al., The liposomal formulation of doxorubicin, Methods in Enzymology, 391,<br />

71–97, 2005.<br />

41. Uster, P. S. et al., Insertion of poly(ethylene glycol) derivatized phospholipid into pre-formed liposomes<br />

results in prolonged in vivo circulation time, FEBS Letters, 386(2,3), 243–246, 1996.<br />

42. Allen, T. M., Mumbengegwi, D. R., and Charrois, G. J. R., Anti-CD19-targeted liposomal doxorubicin<br />

improves the therapeutic efficacy in murine B-cell lymphoma and ameliorates the toxicity of<br />

liposomes with varying drug release rates, Clinical Cancer Research, 11(9), 3567–3573, 2005.<br />

43. Park, J. W. et al., Anti-HER2 immunoliposomes: Enhanced efficacy attributable to targeted delivery,<br />

Clinical Cancer Research, 8(4), 1172–1181, 2002.<br />

44. Gabizon, A., Targeting folate receptor with folate linked to extremities of poly(ethylene glycol)grafted<br />

liposomes: In vitro studies, Bioconjugate Chemistry, 10(2), 289–298, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Folate Receptor-Targeted Liposomes for Cancer Therapy 675<br />

45. Hofland, H. E. J. et al., Folate-targeted gene transfer in vivo, Molecular Therapy, 5(6), 739–744, 2002.<br />

46. Xu, L., Pirollo, K. F., and Chang, E. H., Tumor-targeted p53-gene therapy enhances the efficacy of<br />

conventional chemo/radiotherapy, Journal of Controlled Release: Official Journal of the Controlled<br />

Release Society, 74(1–3), 115–128, 2001.<br />

47. Lee, R. J. and Huang, L., Folate-targeted, anionic liposome-entrapped polylysine-condensed DNA for<br />

tumor cell-specific gene transfer, The Journal of Biological Chemistry, 271(14), 8481–8487, 1996.<br />

48. Gosselin, M. A., Guo, W., and Lee, R. J., Incorporation of reversibly cross-linked polyplexes into<br />

LPDII vectors for gene delivery, Bioconjugate Chemistry, 13(5), 1044–1053, 2002.<br />

49. Pan, X. Q. and Lee, R. J., In vivo antitumor activity of folate receptor-targeted liposomal daunorubicin<br />

in a murine leukemia model, Anticancer Research, 25(1A), 343–346, 2005.<br />

50. Goren, D. et al., Nuclear delivery of doxorubicin via folate-targeted liposomes with <strong>by</strong>pass of multidrug-resistance<br />

efflux pump, Clinical Cancer Research, 6(5), 1949–1957, 2000.<br />

51. Pan, X. Q., Wang, H., and Lee, R. J., Boron delivery to a murine lung carcinoma using folate receptortargeted<br />

liposomes, Anticancer Research, 22(3), 1629–1633, 2002.<br />

52. Qualls, M. M. and Thompson, D. H., Chloroaluminum phthalocyanine tetrasulfonate delivered via<br />

acid-labile diplasmenylcholine-folate liposomes: Intracellular localization and synergistic phototoxicity,<br />

International Journal of Cancer, 93(3), 384–392, 2001.<br />

53. Wang, S. et al., Delivery of antisense oligodeoxyribonucleotides against the human epidermal growth<br />

factor receptor into cultured KB cells with liposomes conjugated to folate via polyethylene glycol,<br />

Proceedings of the National Academy of Sciences of the United States of America, 92(8), 3318–3322,<br />

1995.<br />

54. Reddy, J. A. et al., Folate-targeted, cationic liposome-mediated gene transfer into disseminated<br />

peritoneal tumors, Gene Therapy, 9(22), 1542–1550, 2002.<br />

55. Stevens, P. J., Sekido, M., and Lee, R. J., A folate receptor-targeted lipid nanoparticle formulation for<br />

a lipophilic paclitaxel prodrug, Pharmaceutical Research, 21(12), 2153–2157, 2004.<br />

56. Gabizon, A. et al., In vivo fate of folate-targeted polyethylene-glycol liposomes in tumor-bearing<br />

mice, Clinical Cancer Research, 9(17), 6551–6559, 2003.<br />

57. Turk, M. J., Waters, D. J., and Low, P. S., Folate-conjugated liposomes preferentially target macrophages<br />

associated with ovarian carcinoma, Cancer Letters, 213(2), 165–172, 2004.<br />

58. Pan, X. Q., Wang, H., and Lee, R. J., Antitumor activity of folate receptor-targeted liposomal doxorubicin<br />

in a KB oral carcinoma murine xenograft model, Pharmaceutical Research, 20(3), 417–422,<br />

2003.<br />

59. Guo, W., Lee, T., Sudimack, J., and Lee, R. J., Receptor targeted delivery of liposomes via folate-<br />

PEG-chol, J. Liposome Res., 10, 179–195, 2000.<br />

60. Gabizon, A., Huberty, J., Straubinger, R. M., Price, D. C., and Papahadjopoulos, D., An improved<br />

method for in vivo tracing and imaging liposome using a 67 Gallium-deferoxamine complex, Journal<br />

of Liposome Research, 1, 123–135, 1998.<br />

61. Leamon, C. P., Cooper, S. R., and Hardee, G. E., Folate-liposome-mediated antisense oligodeoxynucleotide<br />

targeting to cancer cells: Evaluation in vitro and in vivo, Bioconjugate Chemistry, 14(4),<br />

738–747, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


34<br />

CONTENTS<br />

Nanoscale Drug Delivery<br />

Vehicles for Solid Tumors:<br />

A New Paradigm for Localized<br />

Drug Delivery Using<br />

Temperature Sensitive<br />

Liposomes<br />

David Needham and Ana Ponce<br />

34.1 Introduction: The Problem and Current Solutions ........................................................... 678<br />

34.2 The Challenges Facing Nanotherapeutic Intervention: A Brief Review<br />

of Lipid-Based Nanocarriers ............................................................................................. 682<br />

34.2.1 Getting the Drug In ............................................................................................. 682<br />

34.2.2 Getting the Nanocapsules to the Tumor ............................................................. 684<br />

34.2.3 Mechanisms of Liposomal Drug Delivery ......................................................... 687<br />

34.3 New Capsular Design for Delivery of Drugs to and Release of Drugs<br />

in Localized Solid Tumors................................................................................................ 691<br />

34.3.1 Getting the Drug Out .......................................................................................... 691<br />

34.3.2 A New Liposome Design.................................................................................... 693<br />

34.3.3 Nanostructures and Membrane Permeability...................................................... 695<br />

34.3.4 In Vivo Growth Delay Studies ........................................................................... 701<br />

34.4 A New Paradigm for Drug Delivery to Tumor Tissues ................................................... 705<br />

34.4.1 Vascular Shutdown ............................................................................................. 705<br />

34.4.2 Exploring Other Tumor Models ......................................................................... 707<br />

34.4.3 Regional Deposition of Drugs Observed Using MRI ........................................ 709<br />

34.4.4 Clinical Trials and Hyperthermia Devices ......................................................... 710<br />

34.5 Conclusions ....................................................................................................................... 711<br />

References .................................................................................................................................... 712<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

677


678<br />

Nanotechnology for Cancer Therapy<br />

34.1 INTRODUCTION: THE PROBLEM AND CURRENT SOLUTIONS<br />

A January <strong>2006</strong> search for the word cancer at www.clinicaltrials.gov yields a list of 4,370 federally<br />

and privately supported clinical trials in human volunteers. Within these results, a focus on<br />

chemotherapy reduces the number to 1,745 (w40% of all trials). 1 Therefore, the discovery and<br />

testing of chemotherapeutic agents for cancer represents a significant effort in current basic and<br />

clinical research. In addition, as cancer is the second leading cause of death in the United States, this<br />

research represents a more than significant investment in health and disease.<br />

Chemotherapeutics are designed to kill cancer cells or prevent them from dividing. The major<br />

types of cancer drugs include antimetabolites, alkylating agents, antibiotics, antimitotics,<br />

hormones, and inorganics (e.g., cisplatin). The National Cancer Institute (NCI) Drug Dictionary<br />

contains technical definitions and synonyms for more than 500 agents that are being used in the<br />

treatment of patients with cancer or cancer-related conditions. 1 However, the magic bullet, * interpreted<br />

here as “a drug that is solely specific for cancer cells,” is still not available. 2,3 The drugs that<br />

have been developed can kill cancer cells, but most of them can also kill normal cells, resulting in a<br />

low therapeutic index (ratio of lethal dose to effective dose). This limits their dosage and, ultimately,<br />

their efficacy. 4 Apart from the success of cisplatin in testicular cancer and methotrexate in<br />

choriocarcinoma, systemic chemotherapy has not yet exposed solid tumors to sufficiently high drug<br />

concentrations for an adequate period of time to cause meaningful tumor regressions. 4 Therefore,<br />

surgery † and radiation are the mainstays for local control of tumors. ‡ In modern clinical practice,<br />

chemotherapy is mostly used as a post-surgical adjuvant to remove microscopic metastases that<br />

have spread from the original site. Other applications include neoadjuvant therapy to downsize<br />

local tumors before surgery or radiation and palliative care for the incurable patient. 4<br />

One solution to the low therapeutic index of these non-specific chemotherapeutic drugs is to<br />

target the drugs to solid tumors <strong>by</strong> associating them with microscopic or submicroscopic carriers<br />

that possess some affinity for the tumors. Microscopic particles such as microspheres accumulate in<br />

tumor capillaries <strong>by</strong> chemoembolism, whereas submicroscopic carriers such as liposomes accumulate<br />

perivascularly through passive extravasation and/or active targeting to cancer cells. 5–10 It is<br />

here, in the context of drug delivery, that nanotechnology has already been explored for over 40<br />

years as an approach for reducing toxicity and increasing the accumulation of drugs in tumors when<br />

compared with free drug. As depicted in Figure 34.1, attention has principally focused on either<br />

self-assembling systems such as polymeric micelles formed from amphiphilic block co-polymers,<br />

11–14 phospholipid-based liposomes, 15,16 and polymer surfactant polymersomes; 17 or<br />

covalent-linked structures such as polymer–drug conjugates, 18–21 dendrimers, 22,23 and radioactive<br />

antibodies. 24,25<br />

The primary goal of nanotechnology for cancer therapy is to change the biodistribution of<br />

drugs, reducing free drug toxicity and favoring tumor accumulation rather than normal tissue<br />

accumulation or excretion. In order to start to achieve this goal, a nanoscopic formulation must<br />

satisfy the well-known requirements of a drug delivery system: load and retain the drug in or on a<br />

submicroscopic or nanoscopic carrier to prevent extravasation in healthy organs and tissues (e.g.,<br />

heart and kidneys for doxorubicin) 26,27 and there<strong>by</strong> reduce toxicity compared to free drug administration;<br />

28,29 evade the body’s defenses (the stealth effect) and avoid premature uptake into the<br />

reticuloendothelial system (RES), making the carrier available in the blood stream; 30 and achieve<br />

carrier uptake in solid tumors <strong>by</strong> the enhanced permeability and retention effect (EPR effect or<br />

*<br />

If an organism is pictured as infected <strong>by</strong> a certain species of bacterium, researchers can hope to discover a drug with a<br />

specific affinity for these bacteria and no affinity for the normal constituents of the body. Such a substance would then be a<br />

magic bullet.<br />

†<br />

About 90% of cancer patients undergo some kind of surgery, but only about 13% of cancers can be cured <strong>by</strong> surgery alone.<br />

‡<br />

Although safety has been improved, people are still dying of the side effects of radiation given for cancer such as Mo<br />

Mowlan, former Northern Ireland Secretary, Aug 19th, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 679<br />

Hydrophilic<br />

Drug<br />

Lipophilic drug<br />

Targeting<br />

antibody<br />

Targeting<br />

group<br />

drug<br />

passive targeting) that is based on the inherent leakiness of tumor microvessels. 31,32 This accumulation<br />

may possibly be aided <strong>by</strong> active targeting using antibodies. 32,33<br />

A recent review <strong>by</strong> Allen and Cullis outlines many of the problems that have been addressed<br />

and solved for nanoscale carrier systems, and it counts the successes for 13 systems that have<br />

gained FDA approval. If the focus is only on the anti-cancer applications, the systems, manufacturers,<br />

indications, and years of approval are: 8<br />

1. The anti-CD20 radioactive antibodies for non-Hodgkin’s lymphoma (NHL), 90 Y-ibritumomab<br />

tiuxetan (Zevalin <strong>by</strong> Biogen IDEC in 2002) 34 and 131 I-tositumomab (Bexxar <strong>by</strong><br />

Corixa and GlaxoSmithKline in 2003).<br />

2. An anti-CD33-linked calicheamicin (Mylotarg <strong>by</strong> Wyeth-Ayerst) for CD33 C relapsed<br />

acute myeloid leukemia (2002).<br />

3. The three liposomal products<br />

a. Stealth (PEG * -stabilized) liposomal doxorubicin (Doxil/Caelyx <strong>by</strong> ALZA and<br />

Schering Plough) approved for Kaposi’s Sarcoma (1995), refractory ovarian cancer<br />

(1999), and refractory breast cancer (Europe 2003 for patients with cardiac risk)<br />

b. Liposomal daunorubicin (DaunoXome <strong>by</strong> NeXstar Pharmaceuticals and Gilead) for<br />

advanced, HIV-associated Kaposi’s Sarcoma (1996)<br />

c. Liposomal doxorubicin (Myocet <strong>by</strong> Elan) for metastatic breast cancer in combination<br />

with cyclophosphamide (Europe 2000)<br />

4. Adjuncts including PEG–granulocyte colony stimulating factor (pegfilgrastim/Neulasta<br />

<strong>by</strong> Amgen) for reduction of febrile neutropenia associated with chemotherapy (2002)<br />

In further depth, how are some of these nanotech-formulated drugs faring in terms of clinical<br />

use and outcome? For the anti-CD20 antibody Zevalin, a randomized, controlled phase III trial<br />

supported its accelerated FDA approval for NHL. 35 This study included 143 subjects with relapsed<br />

or refractory, low grade or follicular NHL, or transformed B-cell NHL. An overall response rate of<br />

80% was obtained in subjects receiving the Zevalin therapeutic regimen (73 subjects) compared to<br />

56% for subjects receiving Rituxan alone (70 subjects). Complete response rates (with no detectable<br />

cancer) were 30% for the Zevalin group and 16% for Rituxan alone. The side effects of Zevalin<br />

drug<br />

drug<br />

drug drug<br />

drug<br />

drug<br />

drug drug<br />

Encapsulated Drug<br />

Targeting <strong>Group</strong><br />

(a) (b) (c) (d) (e)<br />

FIGURE 34.1 Schematic illustrations of the most developed nanoparticles for drug delivery: (a) polymer<br />

micelles; (b) PEGylated liposome showing encapsulated hydrophilic drug, membrane incorporated hydrophobic<br />

drug, and a targeting antibody; (c) a polymer drug conjugate that can also be targeted; (d) a dendrimer with an<br />

encapsulated drug as well as targeting ligands; and (e) a radioactive antibody.<br />

* PEG, Poly(ethylene glycol).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


680<br />

Nanotechnology for Cancer Therapy<br />

include thrombocytopenia and neutropenia. This bone marrow suppression, sometimes persistent in<br />

nature, has led to severe hemorrhaging and infection in a small number of patients. 35,36 Since its<br />

approval in 2002, further studies have supported the efficacy of Zevalin for NHL with response<br />

rates greater than 80%; however, the mean time to progression for these patients has still been<br />

limited to only 9–12 months. 25,37–39<br />

With over 40 years of research dedicated to liposomes alone and over 20,000 papers in the<br />

literature, the question can be asked: has liposomal drug delivery been successful? The first<br />

liposomal carrier, AmBisome (liposomal amphotericin B), was developed for severe fungal infections<br />

40 with a potential market of approximately 100,000 cases per year in the United States.<br />

AmBisome was originally approved in Europe in 1990 (Gilead and NeXstar Pharmaceuticals),<br />

and it was cleared for use in the United States in 1997 (Gilead and Astellas). It is now available in<br />

more than 44 countries worldwide. 41 This formulation then is one of the first successes in therapeutic<br />

nanotechnology. Clinical trials are now underway to examine the expanded use of<br />

AmBisome for prophylaxis against systemic fungal infection following bone marrow transplantation<br />

and other anti-cancer treatments. However, as discussed <strong>by</strong> Becker et al., although the<br />

efficacy of high-dose AmBisome is equal to or better than free amphotericin B, failure rates still<br />

give cause for concern. 42 One explanation for this might be the limited immediate bioavailability of<br />

amphotericin B when it is liposome-encapsulated. This is a good example of a common problem<br />

facing all nanotech drug delivery systems; that is, the challenge of reducing toxicity of free drug is<br />

solved, but the flip side is compromised bioavailability.<br />

As for liposomes against cancer, the FDA has approved two anthracycline formulations,<br />

Daunoxome (liposomal daunorubicin) and Doxil (PEGylated liposomal doxorubicin). These<br />

drugs were originally tested and approved for Kaposi’s Sarcoma, an AIDS-related malignancy<br />

with an incidence of only 0.02%–0.05% in the U.S. 43 Doxil has also recently been indicated for<br />

refractory ovarian cancer that accounts for up to 70% of the 25,500 annual cases of ovarian cancer<br />

in the U.S. 44 Response rates of 12%–20% have been reported for Doxil in this patient group,<br />

proving it to be one of the most effective salvage therapies for ovarian cancer patients, especially<br />

for those that have developed platinum resistance. 45,46 However, complete responses are rare, and<br />

chronic recurrence leads to an average survival time of only 12–24 months after therapy. 44,47 One<br />

of Doxil’s successes is that it is associated with lower incidences of cardiotoxicity, bone marrow<br />

suppression, and nausea than free doxorubicin, but it can cause other side effects, including handfoot<br />

syndrome and stomatitis. 28,48,49 Aside from the formulations currently approved, are new<br />

liposomes in development or are there new indications for previously approved liposomes? A<br />

search for cancer, chemotherapy, and liposome at www.clinicaltrials.gov produced 25 currently<br />

recruiting trials, mainly testing liposomal doxorubicin in combination with other free drugs for a<br />

variety of cancers as listed in Table 34.1. 50<br />

Liposomes were discovered in 1965 51 and explored almost immediately in applications for<br />

drug delivery, principally for cancer. 52 They now represent the most advanced nanoscale drug<br />

delivery system in clinical use. Apart from radioactive antibodies, * other nanoscale molecules and<br />

particles such as polymer–drug conjugates and polymer–drug micelles are being developed but<br />

have not yet progressed to clinical trials. However, even the liposomal nanotechnology effort<br />

accounts for only 25/4370 or 0.6% of all cancer-related clinical trials. This review of the literature<br />

and current clinical trials is illuminating and shows that there is a huge opportunity for nanotechnology<br />

in the form of nanomedicines to impact cancer therapy. If successful (and failed)<br />

designs, tests, and trials are collectively built upon, existing nanotech systems may be integrated<br />

with new nanotech ideas may be integrated to achieve more optimal outcomes for the delivery of<br />

non-specific drugs—that is, a reduced toxicity compared to free drug and an enhanced accumulation<br />

and bioavailability within the tumor. It is here that the thermal sensitive liposome (LTSL)<br />

* A search for monoclonal, antibody, and cancer found 297 trials, but only one phase I, using radio-labeled antibodies.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 681<br />

TABLE 34.1<br />

Cancers Under Investigation for Therapy with Liposomal<br />

Doxorubicin<br />

Breast cancer<br />

Cervical cancer<br />

Endometrial cancer<br />

Male breast cancer<br />

Ovarian epithelial cancer<br />

Ovarian germ cell tumor<br />

Fallopian tube cancer<br />

Peritoneal cavity cancer<br />

Prostate cancer<br />

Non-small cell lung cancer<br />

Small cell lung cancer<br />

Adult hepatocellular carcinoma<br />

Cutaneous T-cell lymphoma<br />

Adult NHL<br />

Mycosis fungoides/sezary syndrome<br />

Adult Hodgkin’s lymphoma<br />

Lymphocyte depleted<br />

Mixed cellularity<br />

Nodular sclerosing<br />

Adult large cell lymphoma<br />

Anaplastic<br />

Recurrent diffuse<br />

Recurrent immunoblastic<br />

Multiple myeloma<br />

offers a new paradigm for local drug delivery and could provide the motivation and experience for<br />

other nanotechnologies that not only encapsulate drugs but also release them, making them<br />

bioavailable at the tumor site.<br />

This chapter will compare and contrast a number of contemporary liposome formulations<br />

(e.g., Doxil, Myocet, DaunoXome, and the currently trialed Onco-TCS). In particular, it will<br />

discuss the rationale and events that have led to the phase I/II testing of the doxorubicincontaining,<br />

temperature-sensitive liposome (ThermoDox) in prostate, breast, and liver cancer.<br />

This a particular example of how old and new concepts, especially ones that are bio-inspired,<br />

may be combined to face the challenges of cancer nanotherapeutics. The new capsular design<br />

addresses an unmet need in liposomal drug delivery, that of getting the drug out of the<br />

nanocarrier and there<strong>by</strong> improving bioavailability. 53 This temperature sensitive liposome formulation<br />

containing lysolipid in the membrane (designated in the literature as LTSL) has been<br />

specifically designed to rapidly release drugs in combination with mild hyperthermic temperatures.<br />

In preclinical trials, the LTSL successfully delivers doxorubicin to solid tumors in<br />

bioavailable quantities that cannot be achieved <strong>by</strong> free drug administration or passive liposome<br />

accumulation, even with stealth liposomes. As depicted in Figure 34.2, this formulation is<br />

establishing a new paradigm for localized drug delivery to tumor tissue—one that moves<br />

nanotechnology for cancer therapy from the requirements of drug retention and carrier extravasation<br />

to the accomplishment of drug release in the blood stream of the tumor, there<strong>by</strong><br />

making the drug bioavailable. This gives the drug access to endothelial cells as well as<br />

neoplastic cells, offering the potential for anti-vascular as well as anti-cancer drug therapies. 54<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


682<br />

FIGURE 34.2 (a) A new paradigm for drug delivery, i.e., drug is released from circulating temperaturetriggered<br />

liposomes in the blood stream of the tumor; drug gains access to tumor endothelia as well as to the<br />

neoplasm. (b and c) Video micrograph images of tumor vasculature before (b) and 24 h after (c) treatment with<br />

temperature-sensitive liposomes containing doxorubicin (Dox–LTSL) with an almost complete elimination of<br />

functioning blood vessels. (From Kong, G., et al., Cancer Res., 60 (24), 6950–6957, 2000; Chen, Q., et al., Mol.<br />

Cancer Ther., 3 (10), 1311–1317, 2004. With permission.)<br />

34.2 THE CHALLENGES FACING NANOTHERAPEUTIC INTERVENTION:<br />

A BRIEF REVIEW OF LIPID-BASED NANOCARRIERS<br />

34.2.1 GETTING THE DRUG IN<br />

Small molecules are on the order of a few nanometers. A liposomal carrier <strong>by</strong> definition must be<br />

larger than the drug it is carrying. Furthermore, a larger liposome can encapsulate a much greater<br />

volume, resulting in a higher drug to lipid ratio (for water-soluble drugs). * This is important<br />

because an overload of carrier material can impair liver metabolism and phagocytosis (RES<br />

uptake) and can cause granulomatous inflammation of the liver. In fact, large doses of empty<br />

liposomes were used in the past to block the RES and promote longer liposome-circulation halflives.<br />

55,56 Figure 34.3 shows that although only two molecules thick (w5 nm), the lipid bilayer †<br />

becomes a significant fraction of the total liposome volume at diameters of 200 nm and smaller. For<br />

reasons to be later discussed, liposomes are traditionally w100 nm in diameter. At this size, the<br />

membrane already takes up 30% of the total liposome volume. Because of mechanical restrictions<br />

imposed <strong>by</strong> membrane thickness and bending stiffness of the lipid bilayer itself (limited curvature),<br />

the smallest possible liposome has a diameter of w30 nm, exhibiting a hydrophobic volume<br />

fraction of w80% and a limited capacity to carry water soluble drugs<br />

Therefore, for these nanoscale carriers, how much drug could be loaded in a 100 nm capsule?<br />

The number of molecules of hydrophilic drug per liposome (N d) depends on the solubility (S d)of<br />

the drug as well as the internal radius (Ri) of the liposome<br />

N d Z 4<br />

3 pR3 i S dA 0<br />

Note that Avogadro’s number is included in the calculation. For cisplatin with a solubility of<br />

5.8 mM at 258C, a 100 nm liposome is capable of encapsulating only 1,340 molecules of this<br />

drug. By comparison, the 100 nm liposome is composed of about w140,000 molecules of lipid,<br />

giving a drug to lipid ratio of only 0.005 (w/w).<br />

The drug doxorubicin, on the other hand, is amphiphilic with a protonatable amine (pKA 8.4); 57<br />

therefore, depending on the pH, a fraction of the drug is soluble in the bilayer. This fraction can<br />

transit the bilayer if an electrochemical gradient is imposed, facilitating the active loading method<br />

*<br />

This chapter does not consider hydrophobic drugs.<br />

†<br />

The hydrophobic bilayer thickness is w4 nm.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Nanoscale Drug Delivery Vehicles for Solid Tumors 683<br />

Volume fraction<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

10<br />

30<br />

50<br />

70<br />

90<br />

110<br />

130<br />

that has revolutionized liposomal drug loading technology for weak acids (Figure 34.4). 58,59 This<br />

method results in extensive loading of up to 61,000 molecules of doxorubicin per liposome (drug to<br />

lipid ratio of 0.3) with the formation of a complex of drug crystals inside the liposome. In considering<br />

new nanoscale drug delivery systems, these kinds of simple calculations are necessary to<br />

determine carrying capacity and drug dosage. As previously mentioned and will be later discussed,<br />

nanoscopic carriers are readily identified <strong>by</strong> the immune system as foreign and can become toxic to<br />

the RES in high doses. There is, therefore, a delicate balance between efficacy and toxicity, not just<br />

for the drug but also for the nanoscopic carrier.<br />

150<br />

170<br />

Outer liposome diameter (nm)<br />

190<br />

Hydrophobic<br />

Aqueous/<br />

Hydrophilic<br />

FIGURE 34.3 Plot of hydrophobic volume fraction as a function of liposome diameter for single bilayer<br />

vesicles. For the traditional 100 nm diameter liposome, 70% of the liposome volume is available for aqueous<br />

content. However, for the smallest liposome possible (30 nm because of limited curvature), there is considerably<br />

more hydrophobic volume per liposome (w80%).<br />

Dox-H + /Citrate fiber<br />

Bundle complex<br />

2<br />

Citrate<br />

Inner surface<br />

outer surface<br />

Citrate<br />

H +<br />

Dox-H +<br />

Dox o<br />

H +<br />

Dox-H +<br />

Dox o<br />

Dox-H +<br />

1<br />

Dox o<br />

H +<br />

Dox-H +<br />

FIGURE 34.4 Remote loading of doxorubicin into liposomes using a pH gradient. Doxorubicin, an amphiphilic<br />

drug, crosses the lipid bilayer in neutral form. When it encounters the acidic interior of the liposome, an<br />

amine is protonated to form Dox-HC. This compound is retained on the interior, forming fiber bundles <strong>by</strong><br />

electrostatic interaction with citrate ions.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Dox o<br />

H +


684<br />

34.2.2 GETTING THE NANOCAPSULES TO THE TUMOR<br />

Nanotechnology for Cancer Therapy<br />

The next issue to address for any nanocarrier is the probability that it will reach the tumor microvessels<br />

after injection into the blood stream. Considerations here include circulation through the<br />

tumor and pharmacokinetics of the carrier. A large fraction of the human cardiac output circulates<br />

through the kidneys and liver, an this explains the rapid clearance of particles <strong>by</strong> these organs. It is<br />

estimated that only a small fraction of the cardiac output actually goes to the peripheral circulation<br />

(and to many tumors). 60 Within this blood, what concentration of liposomal drug is actually<br />

present? Liposome extrusion methods are limited to a lipid concentration of 100 mM, above<br />

which the high viscosity of the lipid suspension limits processing. Assuming a clinical injection<br />

volume of 20 mLs, the lipids would be diluted 250 times (in 5 L blood) to a plasma concentration of<br />

0.4 mM lipid. If drugs such as doxorubicin can be loaded into liposomes at internal drug concentrations<br />

of 300 mM, then the actual plasma drug concentration (still encapsulated) is now only<br />

0.216 mM. This is about 200 times the amount required for a log kill of cells in culture (w1 mM<br />

Dox). However, this assumes complete mixing and no clearance (see below). As the complexities<br />

of drug transport become apparent, it is clear that delivery (and not just reduction of toxicity) is the<br />

major goal of nanotherapeutics.<br />

Given these limits on actual blood levels available for delivery, how does the nanoparticle<br />

accumulate in the tumor? Where does it actually collect within the tumor, and does it stay there?<br />

Either the carrier must lodge in the microvascular bed of the tumor as in chemoembolism strategies,<br />

or it must leak out of the tumor vasculature (<strong>by</strong> the EPR effect) as in most nanoparticle strategies.<br />

The vascular permeability of a drug (across the endothelial lining) is limited <strong>by</strong> the size and charge<br />

of the drug as well as the structures of the vessel wall, particularly the basement membrane. 61 For<br />

small therapeutic drugs such as doxorubicin, permeabilities through normal vasculature are on the<br />

order of 10 K4 cm/s. Similarly, sulforhodamine has been measured to have a permeability of 3.4!<br />

10 K5 cm/s. 62 As the size of a nanoparticle increases beyond 6 nm, its ability to extravasate through<br />

the endothelial lining drops precipitously. 63 For example, the microvascular permeability of<br />

albumin (radius of gyration 7 nm) in normal tissues is 25!10 K8 cm/s, and that for 150,000 MW<br />

dextran (radius of gyration 11 nm) is 7.3!10 K8 cm/s. 62,64–66 Therefore, these polymers and<br />

proteins are well retained in the blood. For 100 nm liposomes, the normal vascular permeability<br />

is also limited at 9!10 K8 cm/s. 65 Therefore, liposomes are over 1,000 times less permeable across<br />

normal vessels than free doxorubicin. It is this physical barrier that is actually responsible for the<br />

reduction in toxicity, dramatically altered biodistribution, and somewhat extended circulation halflife<br />

of liposome-encapsulated drug compared to free drug. For example, conventional liposomal<br />

daunorubicin (Daunoxome) is retained in the blood with a circulation half-life of 4 h compared to<br />

0.8 h for free daunorubicin. 67 As the size of the nanoparticle is increased, motivated <strong>by</strong> the need to<br />

create a large encapsulated volume, the liposome gains the unique ability to keep the drug in the<br />

bloodstream and away from healthy organs and tissues such as the heart and kidneys. This was one<br />

of the major successes of liposomes, a clear demonstration that nanotechnology could impact the<br />

therapeutic toxicity of non-specific drugs like doxorubicin. 28,48,49 In fact, improved toxicity profiles<br />

have been the primary reason for approval of conventional liposome formulations like Daunoxome<br />

(FDA 1996) and Myocet (EU Commission 2000).<br />

But how does this aid in the delivery of drugs to tumors? Fortunately, as already intimated,<br />

tumors have leaky blood vessels that are more permeable than normal blood vessels. For example, the<br />

tumor microvascular permeability for 150 kDa dextran is 5.7!10 K7 cm/s, eight times higher than in<br />

normal tissue, 64 and for albumin, tumor vessel permeability is 25 times higher (1!10 K7 cm/s). 66 It<br />

has been shown that this permeability extends to nanoparticles such as colloidal carbon 68 and liposomes<br />

up to 400 nm in diameter. 69 Therefore, one favored mechanism of targeting tumors has been to<br />

take advantage of the natural leakiness of tumor blood vessels to certain nanoparticles such as<br />

drug–polymer complexes and liposomes. 9,70,71 These drug delivery systems achieve localization<br />

simply <strong>by</strong> passive accumulation. Dosimetry results for the conventional liposome Daunoxome show<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 685<br />

that approximately 2.84 ng/mg (5 mM) of daunorubicin accumulates in the tumor in the first 24 h<br />

after an injected dose of 2 mg/kg body weight. 68 This concentration of drug is actually fairly<br />

successful at killing tumor cells in culture (IC50 2.8 mg/ml). 72 So why does this formulation<br />

show limited success in the clinic? The problem is probably not just inadequate liposome<br />

accumulation but also the inability to make the drug bioavailable in a fast enough time period, as<br />

will be discussed later.<br />

Returning to a discussion of dosimetry, an important challenge was to determine what limits<br />

liposome accumulation in the tumor and how this uptake may be increased. It was soon discovered<br />

that the reason for limited tumor uptake is simply the limited circulation half-life of conventional<br />

liposomes (2–4 h). 73 These liposomes are labeled <strong>by</strong> complement and other opsonins as foreign and<br />

are subsequently removed <strong>by</strong> the phagocytic RES of the liver and spleen. 74,75 Incidentally, particles<br />

larger than 0.6 mm are cleared <strong>by</strong> the spleen regardless of opsonization. 76 This is another major<br />

lesson for any nanotechnology-based drug carrier system—that the surface chemistry of particles<br />

(especially nanoparticles) is crucial in the foreign-body reaction. Essentially, the early researchers<br />

and companies (The Liposome Company/Elan and Nexstar), intending to target cancer, unwittingly<br />

created an artificial parasite. Some formulations, however, have cleverly taken advantage of this<br />

natural localization to the Kupffer cells of the liver for treatment of the parasite Leishmania that<br />

resides in these same cells. 77–79 By creating a highly immunogenic (opsonizable) liposome<br />

containing negatively charged lipids, these formulations were rapidly removed from circulation<br />

<strong>by</strong> the RES. A similar liver-targeting technique was put into clinical practice when Ambisome was<br />

approved for use against leishmaniasis. 80,81<br />

The solution to the clearance problem was to favorably modify the liposome surface chemistry.<br />

Inspired <strong>by</strong> the very long (120 day) half-life of the red blood cell, Allen and Papahadjopoulos 75<br />

explored the idea of incorporating GM1, a red blood cell membrane glycolipid, into the liposomal<br />

membranes (10 mol%). This conferred longevity to the liposomes, increasing plasma half life to<br />

w24 h. However, GM1 is an expensive molecule to purify from whole blood. Therefore, they 82 and<br />

others 83–86 turned to the chemical synthesis and development of polyethylene glycol derivatized lipid<br />

(e.g., DSPE–PEG * ) that similarly increased circulation time. This conjugated lipid had a single<br />

negative charge because the derivatization left the normal zwitterionic PE charged at the phosphate.<br />

Although the incorporation of the negatively charged PS was highly immunogenic, their original<br />

hypothesis was that when incorporated into the bilayer at a few mol%, this shielded negative charge<br />

(similar to that of GM1) was the source of resistance to opsonization (Lasic, D., Personal Communication.<br />

Liposome Technology, Inc., Menlo Park, CA, 1992). Subsequently, evidence has been<br />

accumulated that indicates that the mechanism is primarily one of steric stability because the electrostatic<br />

potential lies somewhat underneath the steric barrier profile provided <strong>by</strong> PEG as demonstrated <strong>by</strong><br />

McIntosh et al. 87,88 The mechanism for the long circulation lifetime conferred <strong>by</strong> incorporating GM1 is<br />

still unknown, although Huang postulated a dysopsonization for this ninja liposome. 89,90<br />

Preclinical animal studies demonstrated the dramatic changes in circulation half-life, tumor<br />

accumulation, and anti-tumor efficacy that resulted from these somewhat bioinspired surface<br />

chemical strategies. 30,82,91–96 The longer circulation time of PEG-grafted lipids allowed for<br />

increased extravasation from leaky tumor microvessels, successfully exhibiting the EPR paradigm<br />

of passive targeting. The new so-called “Stealth” technology was developed <strong>by</strong> Sequus and incorporated<br />

into a PEGylated liposomal doxorubicin formulation, Doxil, <strong>by</strong> Johnson and Johnson.<br />

Doxil has achieved circulation half lives of 20 h in mice 82 and w50 h in humans. 97 Furthermore,<br />

at 48 h after injection, the median tumor concentration of Doxil (in Kaposi’s sarcoma lesions in<br />

humans) was 19 times higher than in normal skin. 97 In mice, about 7 ng/mg of Dox accumulated in<br />

FaDu xenografts (human squamous cell carcinoma) after injection of 5 mg/kg Doxil<br />

(Figure 34.24). 98<br />

* DSPE–PEG, Distearoylphosphatidylcholine grafted with polyethylene glycol.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


686<br />

Expansion modulus K A mN/m<br />

4000<br />

3500<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

0 20 40 60 80 100 120 140<br />

t 1/2 (min)<br />

FIGURE 34.5 Plasma half life of conventional liposomes versus elastic modulus of the membrane (for normal<br />

RES). Membranes with a higher elastic modulus circulate for longer times. That is, a low resistance to<br />

mechanical expansion of the bilayer interface is correlated with liposome destruction in the blood stream.<br />

This points to a mechanical mechanism for the presumed opsonization of liposomes. A similar trend holds<br />

when the RES is pre-saturated with liposomes, but half-lives are on the order of hours.<br />

An alternative mechanism to resist opsonization that also relies on reducing protein adsorption<br />

and binding at the bilayer head group region was to modify the membrane mechanically rather than<br />

to chemically modify the surface. This was achieved <strong>by</strong> using a lipid composition that creates a very<br />

tight interface among the headgroups as measured <strong>by</strong> a high area dilation modulus for the bilayer. 99<br />

This composition utilizes lipids with saturated acyl chains and incorporates high concentrations<br />

(w50 mol%) of cholesterol in the membranes. Taking data from the literature on circulation half<br />

lives and comparing them to the elastic area dilation modulus measured using the micropipette<br />

technique for similar lipid compositions, a direct correlation was found between circulation half life<br />

and bilayer elasticity as shown in Figure 34.5. 87<br />

This shows that resistance to bilayer area expansion is a key property in achieving a long<br />

circulation performance for these nanocarriers. That is, a high resistance to mechanical expansion<br />

of the bilayer interface is correlated with reduced opsonization in the blood stream. This points to a<br />

nanomechanical mechanism for the presumed opsonization of liposomes where the molecular<br />

insertion that underlies opsonization must create area expansion in the bilayer. A similar trend<br />

holds when the RES is pre-saturated with liposomes. Here, half-lives are on the order of hours, and<br />

clearance may be due to destruction in the blood stream. Again, low membrane compliance resists<br />

this and is the property that correlates with long circulation performance.<br />

A particular example is the sphingomyelin/cholesterol bilayer with a high area dilation<br />

modulus of w2,200 mN/m 99 that has been utilized in Onco-TCS, the liposomal vincristine *<br />

formulation of Inex. 100 Onco-TCS achieves a prolonged circulation half life (6.6 h) when<br />

compared to free vincristine (1.36 h) in preclinical studies. 101 Therefore, a low bilayer membrane<br />

compliance can be used to increase circulation time. Onco-TCS also exhibits a higher concentration<br />

in tumors and demonstrates less neurotoxicity and greater efficacy in mice bearing L-1210<br />

and P-388 leukemias. This formulation has been shown to have a similar MTD when compared<br />

with free vincristine, and it is currently in phase II/III clinical trials for aggressive recurrent<br />

NHL. 8,102–104<br />

* First extracted from the Madagascar periwinkle.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Nanoscale Drug Delivery Vehicles for Solid Tumors 687<br />

34.2.3 MECHANISMS OF LIPOSOMAL DRUG DELIVERY<br />

As the study of liposomes continued, it was clear that in order to address mechanistic questions<br />

about drug delivery such as the EPR effect and the distribution of liposomes and other nanoparticles<br />

within the tumor, a direct in vivo measurement in real time was required. The skin flap tumor<br />

window chamber, first used in the 1940s for observation of tumor microvasculature, offered this<br />

capacity. 105 This research team 54,71,106 and others 9,66 have extensively utilized the window<br />

chamber to measure liposomal circulation half-life, accumulation in tumors, and tumor distribution<br />

both with and without hyperthermia. For the conventional, stealth, and temperature-sensitive liposomes,<br />

this technique has been invaluable in the direct observation of accumulation and/or<br />

triggered release of encapsulated drug in the tumor tissue. The window chamber is shown in<br />

Figure 34.6.<br />

The procedure involves implanting tumor cells onto the exposed fascia and covering the<br />

surgical site with a glass window. After the tumor grows, the vasculature is visible <strong>by</strong> intravital<br />

microscopy with transmitted white light. In addition, using epifluorescence, it is possible to quantitatively<br />

evaluate fluorescent or fluorescently labeled molecules within the vasculature and within<br />

the tumor tissue over time. With this model, complex drug delivery data such as pharmacokinetics,<br />

vessel wall permeability, and intratumoral distribution have been quantified. 105 Using small fluorescent<br />

molecules fluorescently labeled albumin, 62 and fluorescently labeled liposomes, 65 the<br />

following have been shown:<br />

1. Small molecules such as sulforhodamine (606 Da) and proteins such as albumin<br />

(66,430 Da) are readily extravasated into tumor tissue, supporting the established permeability<br />

of tumor vasculature. 62<br />

2. Conventional liposomes also extravasate but the extent of accumulation in the perivascular<br />

space is severely limited <strong>by</strong> fast RES uptake and disappearance from circulation<br />

(Figure 34.7, middle panels and right graph). 65<br />

3. Stealth liposomes have an extended circulation half-life and accumulate to a greater<br />

extent as shown in the epifluorescent images of the tumor vasculature in Figure 34.7<br />

(left panel and left graph). 65 This supports the early efficacy studies <strong>by</strong> Vaage and<br />

FIGURE 34.6 (See color insert following page 522.) Schematic of rat with a window chamber surgically<br />

attached to the dorsal skin flap (left). The skin flap is a 200 mm thick section of dermis with implanted tumor<br />

cells. As these cells grow, the tumor is visible via trans-illumination (right) or epi-illumination. This model was<br />

used to quantify complex data such as tumor vascular permeability and liposome accumulation. (From Huang,<br />

Q., et al., Nat. Biotechnol., 17, 1033–1035, 1999. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


688<br />

(x) 90 min after injection' 1 min after injection Transillumination<br />

(a) (b)<br />

100 DOXIL-Like<br />

100<br />

Normalized liposome amount<br />

80<br />

60<br />

40<br />

20<br />

Stealth in tumor<br />

0 0 15 30 45 60 75 90<br />

Time (min)<br />

Conventional in tumor<br />

(a) (b) (c)<br />

(d) (e) (f)<br />

(g) (h) (i)<br />

Vascular<br />

Interstitial<br />

Normalized liposome amount<br />

Metcalfe that showed that stealth, but not conventional, liposomes produced tumor<br />

efficacy in mouse tumor models. 93–96<br />

What this data 99,107,108 clearly shows is that the stealth liposomes out-perform the<br />

conventional liposomes with respect to circulation half-life and accumulation in<br />

the heterogeneously leaky tumor (left and center panels). Furthermore, as shown in the<br />

right panel, normal blood vessels (in the same window chamber) maintain a seal against<br />

liposome extravasation. The small 100 nm diameter of the liposome and the incorporation<br />

of PEG-lipid (4–5 mol%) work to preferentially increase the amount of encapsulated<br />

80<br />

60<br />

40<br />

20<br />

Nanotechnology for Cancer Therapy<br />

Stealth in normal<br />

EVACET-Like<br />

0 0 15 30 45 60 75 90<br />

Time (min)<br />

Vascular<br />

Interstitial<br />

FIGURE 34.7 Video micrographs showing the window chamber vasculature via trans-illumination (top row)<br />

and epi-illumination (second and third rows) after injection of stealth liposomes in tumor (left column),<br />

conventional liposomes in tumor (middle), or stealth liposomes in normal tissue (right). The second row<br />

shows 1 min after injection, and the third row shows 90 min after injection. The graphs show the relative<br />

amounts of liposomes in the blood stream and in the tumor versus time (stealth on left and conventional on<br />

right) as calculated from the fluorescent images. (From Wu, N. Z., et al., Cancer Res., 53 (16), 3765–3770,<br />

1993. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 689<br />

FIGURE 34.8 Schematic showing the traditional paradigm for liposomal (and polymer) drug delivery. In this<br />

paradigm, extended circulation time takes advantage of leaky tumor vasculature to preferentially deliver<br />

carriers to the tumor tissue.<br />

drug being delivered to the tumor while normal tissue is protected. This behavior<br />

establishes the current paradigm of the EPR effect for liposomes (and other macromolecules<br />

and nanoparticles) as depicted in Figure 34.8; that is, the basis for<br />

preferential drug delivery is this disparity in vascular permeability to even relatively<br />

large 100 nm particles, resulting in a fraction of the injected dose gaining access to the<br />

perivascular space of the tumor tissue.<br />

4. The vascular pore cut-off size for liposome extravasation in the human ovarian carcinoma<br />

SKOV3 was found to be between 7 and 100 nm diameter at 378C; 71 other studies have<br />

shown a cut-off size of up to 400 nm in the human colon adenocarcinoma LS174T. 70<br />

5. Mild local hyperthermia (HT) of 428C could dramatically increase the permeability of<br />

tumor vessels to stealth (non-thermal sensitive) liposomes up to 400 nm in diameter with<br />

the most dramatic effect seen at 100 nm. 71 In one experiment, the amount of liposome<br />

extravasation was increased <strong>by</strong> 50-fold with HT compared to normothermic controls for<br />

the relatively impermeable SKOV3 tumor. 109<br />

6. In spontaneous tumors in pet cats, HT dramatically increased the accumulation of nonthermal<br />

sensitive technetium-99-labeled liposomes in the tumor tissue <strong>by</strong> 2–16-fold,<br />

further supporting the utility of HT in spontaneous tumors. 106<br />

These results demonstrate the important benefits of mild hyperthermia for liposome distribution<br />

in the EPR paradigm (even with non-thermal sensitive liposomes). However, they also point to a<br />

number of flaws in relying on the EPR effect alone to deliver nanoparticles and their drug. It is<br />

apparent that not all tumors are permeable to nanoparticles on the scale of 100 nm. In addition,<br />

throughout the window chamber studies, researchers have observed that liposomes and other<br />

nanoparticles only accumulate in the perivascular space after extravasation. 9,71,109 Even<br />

antibody-targeted liposomes show only limited interstitial diffusion.<br />

As for the visualization of liposome distribution, Figure 34.9 shows magnified videomicrograph<br />

images of the microvasculature of a mammary adenocarcinoma model in a skin flap window<br />

chamber. 65,110 On the left, a trans-illuminated image shows blood vessels of approximately 30 mm<br />

in diameter. The smallest (arrow) is about 10 mm, allowing only slow passage of red cells in single<br />

file. On the right is an epi-illuminated image of fluorescent liposome distribution 90 min after<br />

injection of NBD-labeled liposomes. This characteristic hot spot pattern illustrates the gross heterogeneity<br />

of tumor vascular permeability and the limited transport of fluorescently labeled liposomes<br />

into the packed cellular bed of the tumor tissue. This image very dramatically demonstrates that<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


690<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 34.9 (a) Videomicrograph bright-field image of the microvasculature of a mammary adenocarcinoma<br />

in a skin flap window chamber. (b) Videomicrograph epifluorescent image of the same window chamber,<br />

showing a hot spot pattern of stealth liposomes 90 min after tail vein injection. The focal deposits of liposomes<br />

are limited to the perivascular space. (From Chen, Q., et al., Mol. Cancer Ther., 3 (10), 1311–1317, 2004.<br />

With permission.)<br />

liposomes only accumulate in the perivascular space of the tumor and do not diffuse to cells deep<br />

within the tumor. In fact, the hot spots are close enough to the circulation that a portion of the<br />

liposomes can actually wash back into the vessels. 111 The important conclusion here is that the<br />

liposome-encapsulated drug does not reach the majority of the tumor cells based on the EPR effect<br />

alone. This is also expected for other nanoparticles that aim to deliver drugs to tumors; they will not<br />

reach beyond the perivascular space unless some additional chemical or external influence<br />

is exerted.<br />

Whereas distribution of nanoscopic particles only to the perivascular space is clearly a limitation,<br />

a bigger question is if the drug is actually bioavailable when delivered to even these regions.<br />

How does the drug get to its target once the carrier has accumulated in the tumor? How do the<br />

properties of the carrier and the local tumor environment contribute to this process? This chapter<br />

previously described how the currently approved liposomal nanotechnologies achieve a significant<br />

decrease in drug toxicity and a modest accumulation of drug in the tumor. However, the design<br />

requirement of retaining drug in the carrier during circulation results in most of the drug remaining<br />

inside the carrier if and when it localizes in the tumor tissue. It is expected that the drug will<br />

passively leak across the membrane, but it seems that it is often not bioavailable in high concentrations<br />

on the time scale necessary for therapeutic action. In particular, non-cell cycle specific<br />

drugs like doxorubicin should ideally be available to the intracellular targets (including DNA and<br />

RNA) as a sharp peak of concentration, maximizing drug effect in a short time period to prevent cell<br />

recovery and resistance. This inherent slow release may actually be an advantage for drugs that<br />

have cell cycle specificity such as the vincristine formulation Onco-TCS. 104 In this scenario, the<br />

slow release and long circulation half life of the sphingomyelin/cholesterol liposomes make free<br />

vincristine available to as many cancer cells as possible as they progress through the cell cycle to<br />

the G2/M stage.<br />

Overall, the controlled release of the drug from carriers at the tumor site is a problem that all<br />

polymer, liposomal, micro-particle, and capsular nanotechnologies still face. 8 For polymer–drug<br />

conjugates, it has been fairly straightforward to convert the covalent linkers into biodegradable<br />

linkers (cleaved <strong>by</strong> endogenous enzyme action or <strong>by</strong> simple bond hydrolysis). Preclinical studies<br />

using this method have demonstrated drug release specifically at the tumor with a resulting increase<br />

in anti-tumor efficacy. 112,113 However, for liposomes, as mentioned above, the main focus of<br />

modern carrier design has been to lessen leakage from the liposomes in the circulation phase<br />

rather than to trigger release at the tumor site.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 691<br />

34.3 NEW CAPSULAR DESIGN FOR DELIVERY OF DRUGS TO AND RELEASE<br />

OF DRUGS IN LOCALIZED SOLID TUMORS<br />

This section will briefly review the events leading up to the invention of the temperature sensitive<br />

liposome (LTSL), a potential solution to this paradox of drug retention in circulation and rapid<br />

release at the tumor site. It will also discuss studies on the composition, structure, and property<br />

relationships of membranes. It was this wealth of fundamental lipid bilayer knowledge that allowed<br />

a liposome to be forward engineered to perform a number of functions: survive the physical and<br />

chemical conditions of the blood stream, accumulate in the tumor, and respond to stimuli (in this<br />

case, mild hyperthermia) to trigger the release of encapsulated drug. The key to the design and its<br />

functional performance, as in all nanotechnologies for drug delivery, is to understand the materials<br />

relationships of the carrier.<br />

34.3.1 GETTING THE DRUG OUT<br />

After 35 years of research, the long-circulating liposomes (e.g., Doxil) have fulfilled the requirements<br />

of loading and retention of drug, evasion of the body’s defenses, and accumulation in tumors<br />

<strong>by</strong> EPR. However, it is widely recognized that liposome performance is often hampered <strong>by</strong> the very<br />

lipid compositions that create low bilayer permeability and allow retention of the encapsulated drug<br />

during the blood-borne transport phase. The remaining goal of releasing a drug such as doxorubicin<br />

fast enough to kill tumor cells was not part of the original conventional (Myocet, Daunoxome) or<br />

stealth (Doxil) designs. One solution that now comprises the main topic for the rest of this chapter<br />

has been to take a materials engineering approach to liposome design and to explore nanoscale<br />

material defects in these 100 nm carriers. The task undertaken for this particular system was to<br />

characterize the relationships between lipid composition (lipids and lysolipids), bilayer nanostructure<br />

(grain boundary defects), and membrane properties (mechanical, thermal, and permeability) in<br />

order to design a liposome that releases drug rapidly upon an externally applied trigger, in this case,<br />

mild heat of around 428C. 114,115<br />

The functions (load, retain, target, and release) have been defined, and there is an established<br />

wealth of materials data, prototype tests, production methods, and performance studies for liposomes<br />

in the literature. Using these guidelines, the goal is to trigger release of the liposomes after<br />

extravasation <strong>by</strong> EPR. Based on size alone, liposomes greater than 600 nm are taken up <strong>by</strong> the<br />

spleen, and those less than 70 nm are taken up <strong>by</strong> hepatocytes. 76,116 In addition, 400 nm is the<br />

cut-off for tumor extravasation. 70 Therefore, the optimal liposome size for EPR-based drug delivery<br />

is about 100 nm. In addition, 100 nm liposomes have demonstrated the greatest increase in tumor<br />

uptake when combined with hyperthermia. 109 As for surface design, conventional formulations<br />

only provide a relatively short circulation half life (2–4 h) 73 , but PEGylated lipids with their surface<br />

repulsive potentials 87 tend to avoid RES uptake. Therefore, the starting point was the 100 nm<br />

stealth formulation, shown schematically in Figure 34.10.<br />

Given this liposome, researchers must remember that many physical limitations still hamper<br />

effective drug delivery. As discussed above, only a small fraction of the circulation passes through<br />

the tumor. Some tumors have low permeability, preventing uptake of even 100 nm nanoparticles. 71<br />

Upon extravasation, liposomes accumulate only in the perivascular space, and the liposomes that do<br />

accumulate achieve only slow release of bioavailable drug. In 1995, even though liposomal formulations<br />

were being tested and approved, the current technology was not achieving the kind of<br />

bioavailable drug levels necessary for increased efficacy. Through collaboration with the<br />

hyperthermia program at Duke Medical Center, this team found that local hyperthermia may be<br />

used to address many of these problems. Before discussing the materials aspects of the liposome<br />

that promote triggered release, the effects of hyperthermia on liposomal drug delivery must<br />

be considered.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


692<br />

Drug<br />

FIGURE 34.10 A tentative liposome design was a 100 nm diameter stealth liposome composed of a DPPC<br />

bilayer with 4–5 mol% PEG–DSPE (only shown on the outside) and encapsulating the drug doxorubicin via<br />

remote loading.<br />

The mechanisms of synergy between hyperthermia (HT) and liposomal chemotherapy are<br />

manifold (Figure 34.11). HT improves vascular perfusion, there<strong>by</strong> increasing blood flow through<br />

tumors that usually have significant areas of poor perfusion. Furthermore, HT dramatically<br />

enhances transvascular permeability, increasing pore size between endothelial cells to enhance<br />

liposome extravasation. Tumor microvessels are particularly sensitive to the effects of HT,<br />

providing added selectivity that protects the normal tissues. 117,118 These changes last for several<br />

hours following HT treatment and have been exploited to augment liposomal drug delivery to<br />

tumors up to 50-fold in otherwise impermeable tumor vessels as described above. 109 In addition,<br />

hyperthermia provides direct cytotoxicity to tumor cells and synergism with many drugs, including<br />

doxorubicin and cisplatin. 117,119 A number of liposomal drugs have been tested in a variety of<br />

Direct<br />

cytotoxic<br />

effects<br />

Hyperthermia<br />

Synergistic<br />

effects with<br />

DOX<br />

Increased<br />

therapeutic effect<br />

Thermosensitive<br />

liposomes<br />

Increased<br />

extravasation<br />

Increased<br />

local drugh<br />

concentration<br />

Nanotechnology for Cancer Therapy<br />

Triggered<br />

drug release<br />

FIGURE 34.11 Schematic showing the mechanisms of synergism between hyperthermia and liposomal drug<br />

delivery. HT alone has some direct cytotoxic effects as well as synergism with many chemotherapeutic agents.<br />

In addition, HT enhances liposome extravasation and local tumor perfusion. The new concept was rapid<br />

triggered drug release from liposomes at mild hyperthermic temperatures.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 693<br />

tumor models, and these reports consistently prove that HT increases liposome accumulation in<br />

tumors and enhances anti-tumor effect. 117 Finally, with new advances, the current HT technology<br />

has the ability to administer controlled local heat to superficial or deep locations with accurate<br />

thermal dosimetry. 118 The remaining question, then, is if the performance of stealth liposomes can<br />

be improved <strong>by</strong> releasing the drug using externally applied heat. The answer lies in tailoring the<br />

materials composition, nanostructure, and properties of the lipid bilayer itself.<br />

It is important to note here that the traditional target for chemotherapeutic drug delivery has<br />

always been distant metastases; very few primary tumors have been treated, much less cured, with<br />

chemotherapy alone. Therefore, these studies deviated from the traditional target for liposomes, took<br />

a bold and possibly unconventional track, and focused on local tumors. Whereas metastases are<br />

clearly the most important problem, an inability to control the local growth of solid tumors leads to<br />

significant morbidity and mortality in cancer patients. In the U.S., over 60% of therapies eventually<br />

fail with local recurrence. 120 Furthermore, improved locoregional control has been shown to enhance<br />

survival in many common cancers, including breast, prostate, lung, and cervical cancer. 121,122 What<br />

is needed here is a parenteral drug delivery system that can dramatically reduce the tumor size, to<br />

debulk for subsequent surgery or radiation, or maybe even achieve complete regression.<br />

What this team set out to do was to design, create, and test a liposome that would respond to<br />

clinically attainable mild hyperthermia temperatures in the range of 418C–428C. The design<br />

required an intravenously injected liposome that released its drug only upon entering the heated<br />

tumor mass. The great surprise, in the very first series of preclinical tumor growth delay studies,<br />

was a response of 11 out of 11 complete regressions. 123 In addition, measurements of drug accumulation<br />

in the tumor tissue showed concentrations three times higher than with the combination of<br />

Doxil-like liposomes and HT 98 in spite of similar plasma half lives. *<br />

34.3.2 A NEW LIPOSOME DESIGN<br />

The lysolipid-containing temperature-sensitive liposome (LTSL) was invented in 1996, offering an<br />

effective way to achieve targeted tumor drug release for cancer chemotherapy. 98,114 This formulation<br />

is now in phase 1 human clinical trials. 53 Preclinical studies in mice, rats, and dogs have<br />

shown promising efficacy, reducing tumor growth in many of the tumors tested. 50,98,123 However,<br />

some variations in tumor response and subtleties in mechanism are still to be investigated. As will<br />

be discussed at the end of this section, attempts to correlate tumor growth delay with measurable<br />

parameters such as hypoxic fraction, intratumoral drug concentration, and tumor drug sensitivity<br />

reveal a complex relationship between these parameters. The current discourse will focus on how<br />

this liposome achieves the performance requirement of rapid triggered release of doxorubicin at<br />

clinically attainable, mild hyperthermic temperatures.<br />

As demonstrated w35 years ago <strong>by</strong> de Kruijff 124 and Papahadjopoulos 125 and subsequently<br />

modeled <strong>by</strong> Mouritsen, 126 the permeability of a lipid bilayer membrane to small ions, water, and<br />

even small drug molecules is dramatically increased at its gel-to-liquid melting phase transition<br />

when compared to its gel state. At the melting phase transition temperature (Tm), gel and liquid<br />

domains coexist, and solid phase lipids form grain structures to be described below. At the grain<br />

boundaries, leaky interfacial regions are present. 126–129 The increased permeability may be attributed<br />

to these interfacial regions as well as the much higher lateral compressibility of the membrane<br />

in the phase transition region. 130,131 For HT-triggered drug release in the body, a lipid was needed<br />

that remained solid and impermeable at body temperature (378C) but became leaky to its encapsulated<br />

drug when heated just above this temperature. The lipid of choice was relatively obvious,<br />

DPPC † has a phase transition temperature of 41.98C 132 and so has been used in many studies that<br />

aimed to explore temperature-sensitive drug delivery. 117,133<br />

*<br />

The LTSL, if anything, showed a shorter half-life.<br />

†<br />

DPPC, Dipalmitoylphosphatidylcholine.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


694<br />

Thermal sensitive liposomes (TSLs) were not new in 1996. Previous formulations composed of<br />

DPPC mixed with other lipids like DSPC * , 133 DPPG † , 134 or HSPC ‡ :Cholesterol, 135 all required a<br />

temperature of 438C–458C to reach the peak in membrane permeability. The two- (or more)<br />

component phase diagrams show the origin of this transition temperature elevation, an ideal<br />

thermodynamics of mixing (entropic) that broadens the transition and raises it compared to the<br />

pure DPPC bilayer. 136–138 As a result, these formulations were not well matched to the clinical<br />

setting as hyperthermia temperatures greater than 438C are not clinically attainable and may cause<br />

discomfort and vascular damage to the patient. 118 Also, even at the transition temperature, the rates<br />

of drug release for these lipid mixtures were still relatively slow (15–60 min). This is much longer<br />

than the transit time of a liposome through the tumor circulation, so these early TSLs relied on the<br />

EPR effect to position the liposomes in the tumor before releasing the drug. Unfortunately, the<br />

stealth technology had not yet been invented when these TSLs were being tested, and they did not<br />

accumulate in the tumor enough to be efficacious. Paradoxically, the stealth liposomes achieved<br />

accumulation but did not release their drug. Because many development agendas for conventional<br />

and stealth liposomes were already in place, TSLs received less attention for many years. With a<br />

renewed interest in thermally triggered systems and with the benefit of hindsight, this team began<br />

the LTSL design in 1996 <strong>by</strong> combining old and new technologies (thermal sensitivity and stealth<br />

PEG). In addition, experience from many years of studying, measuring, and modeling the material<br />

properties of lipid bilayers using micropipette manipulation techniques was applied.<br />

The micropipette manipulation system has been developed and used <strong>by</strong> researchers to measure<br />

the mechanical properties of lipid bilayer vesicles, 99,107,108 permeability of lipid membranes, 139–141<br />

molecular exchange with lysolipids, 142 and other mechanical and thermomechanical properties<br />

143,144 in order to characterize bimolecular membranes as previously reviewed. 145–147 The<br />

micromechanical experiments on lipid vesicles established several key findings that influenced<br />

the new design.<br />

1. Experiments that measured the resistance of lipid vesicle bilayers to area expansion and<br />

so derived the bilayer elastic expansion modulus showed that liquid state (La) lipids such<br />

as egg PCs plus cholesterol provided sufficient strength to retain drug and remain intact<br />

in the blood stream.<br />

2. The more saturated PCs § and sphingomyelin plus cholesterol exhibited enormous elastic<br />

moduli approaching traditional plastics 99 and were even more impermeable to water 139<br />

and especially drugs.<br />

3. The inclusion of cholesterol (to achieve strength and impermeability) abolished the<br />

melting phase transition and so attenuated any possible thermal sensitivity for DPPC<br />

(e.g., DPPC:Cholesterol in a molar ratio of 1:1). 135<br />

4. Even non-cholesterol-containing PC bilayers had moderate strength as gel phase<br />

membranes that were essentially thin (5 nm) layers of wax. 107<br />

Therefore, a liposome using gel phase lipids was designed. The design was built around DPPC<br />

with its T m of 41.98C plus a PEG -conjugated lipid (DSPE–PEG). The inclusion of DSPE–PEG did<br />

not change the transition permeability compared to pure DPPC. 148 Ideally, a liposome with a<br />

slightly lower Tm and higher membrane permeability than that of pure DPPC is desired. The<br />

additional component that achieved these goals will be discussed later in this section. First, the<br />

*<br />

DSPC, Distearoylphosphatidylcholine.<br />

†<br />

DPPG, Dipalmitoylphosphatidylglycerol.<br />

‡<br />

HSPC, Hydrogenated Soy Phosphatidylcholine.<br />

§ Phosphatidylcholines.<br />

Poly(ethylene glycol)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Nanoscale Drug Delivery Vehicles for Solid Tumors 695<br />

structural features that apparently made the liposomes leak their drug, i.e., the grains and grain<br />

boundaries, must be discussed.<br />

34.3.3 NANOSTRUCTURES AND MEMBRANE PERMEABILITY<br />

In 1987, working with giant lipid vesicles (20–30 mm diameter), this team showed that a single lipid<br />

vesicle could be heated or cooled through its main phase transition at the tip of a micropipette. 107,143<br />

This allowed the transition to be characterized in terms of the bilayer area change reflecting a<br />

w25% change in cumulative area per molecule (from 65A 2 in the liquid phase to w41A 2 in the Pb 0<br />

solid phase). Under no applied membrane tension, the solid vesicle was crumpled and faceted,<br />

whereas under controlled low tension (suction), it formed flat plate-like domains. Upon melting,<br />

these lipid 2D icebergs flowed past the pipet tip as they moved in a sea of melting lipid. In the gel<br />

(solid) state, it was shown that the rippled nanosuperstructure that characterized the P b 0 phase could<br />

be pulled flat. 143 Then, subjecting the vesicle to shear deformation, a yield shear and shear viscosity<br />

for the membrane were measured. The values for these material properties were on the order of<br />

simple waxes and plastics just below their melting temperatures. The membrane behaved like a<br />

Bingham plastic, yielding after an elastic shear deformation and then flowing with a velocity that<br />

was proportional to the excess stress (provided <strong>by</strong> the micropipette suction pressure) above the<br />

yield point. Both the yield shear and shear viscosity increased with decreasing temperature below<br />

the lipid melting temperature, showing that the more solid a membrane, the more difficult it was to<br />

deform in shear. It has been established that, for crystalline materials in general, the yield shear and<br />

shear viscosity reflect density and mobility of crystal defects in the solid structure. In 1987, this<br />

team postulated that grain boundaries and/or intra-grain dislocations allowed the shear deformation<br />

for the 2-molecule thick wax membrane.<br />

It was expected that because crystalline solids nucleate in liquids upon cooling, the solid lipid<br />

membrane should form as grains (analogous to panels in a soccer ball; see Figure 34.12) that meet<br />

at grain boundaries (stitches in the soccer ball). This grain-faceted nanostructure was observed in<br />

subsequent electron micrograph images of the 100 nm diameter liposome formulation containing<br />

65 2<br />

A<br />

Liquid<br />

(liquid<br />

crystalline)<br />

Waxy lipid membrane<br />

2 molecules (4 nm) thick<br />

40 2 A<br />

40A<br />

Tc 54A<br />

Solid<br />

(gel)<br />

100 nm in diameter<br />

100 nm<br />

FIGURE 34.12 Schematic showing the soccer ball-like structure of the waxy liposome (center). The electron<br />

micrograph images show solid-phase liposomes, 100 nm in diameter, that are loaded with doxorubicin. The<br />

nanoscale grains and grain boundaries are evident from the faceted appearance. The phoshpholipid bilayer is 2<br />

molecules thick (w4 nm), and there is a difference in molecular area and thickness between solid and liquid<br />

phase lipids (left). (From Ipsen, J. H. and O. G. Mouritsen, Biochim. Biophys. Acta, 944 (2), 121–134, 1988.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


696<br />

(a) (b) (c)<br />

5 μm<br />

1 μm<br />

40 μm<br />

FIGURE 34.13 Solid lipid monolayers on gas microbubbles also show grain structure as revealed <strong>by</strong> freeze<br />

fracture electron microscopy (a and b, 5 mm and 1 mm scale). (c) Fluorescence micrograph images show that<br />

Bodipy lipids included in the lipid formulation at w1 mol% are excluded from the solid grains and forced<br />

to accumulate at the grain boundaries. (From von Dreele, P. H., Biochemistry, 17 (19), 3939–3943, 1978.<br />

With permission.)<br />

doxorubicin (Figure 34.12). 149 It was confirmed that this grain structure was evident at the nanoscale<br />

(100 nm) and for the 2-molecule thick membrane. Also shown in Figure 34.12 is a schematic of<br />

the solid/liquid interface, illustrating the mismatch in lipid molecular area and thickness between<br />

solid and liquid states. It is this nanoscale defect that provides the anomalous membrane permeability<br />

that has been observed and modeled in the transition region 125,126,150 and that is the<br />

underlying key to the LTSL drug release mechanism.<br />

Additional evidence for the grain structure came from micromechanical and ultrastructural<br />

studies of solid lipid monolayers at microgas bubble interfaces (using similar saturated PC<br />

lipids). This work aimed to create and characterize lipid-stabilized gas-filled ultrasound contrast<br />

agents. 151 It was observed that the yield shear and shear viscosity were consistent with a polycrystalline<br />

monolayer, and both freeze fracture electron microscopy and fluorescence microscopy<br />

showed the grain structure in these micro-lipid shells (Figure 34.13).<br />

With the grain structure now characterized, the goal of enhancing the permeability at the phase<br />

transition (compared to pure DPPC and other lipid mixtures) was now the priority. A series of<br />

micropipette experiments measured and modeled the exchange of molecules that entered the<br />

bilayer interface (tannic acid) 152 or incorporated into or even crossed the bilayer (peptides 153<br />

and lysolipids) 142,154 and could then be washed out with bathing media. By measuring the area<br />

change of a single vesicle and <strong>by</strong> knowing the areas per molecule of the SOPC * bilayer lipid (65A 2 )<br />

and the MOPC † lysolipid (50A 2 ), the mol fraction of lysolipid entering the bilayer from the aqueous<br />

solution was measured, down to a fraction of a mol%, was measured. Importantly, the desorption of<br />

lysolipid upon wash-out with lysolipid-free media was also measured, finding that it leaves the<br />

outer monolayer rapidly (within seconds) 142 . This team then discussed the possibility that the<br />

membrane permeability to small drugs might be increased if a certain amount of water-soluble<br />

lysolipids were incorporated in the solid bilayer and then desorbed from the membrane upon<br />

heating through the transition. The lysolipid MPPC ‡ is structurally and chemically compatible<br />

with the bilayer lipid DPPC § , but it is quite soluble in water, especially as micelles. It was expected<br />

that MPPC would leave the bilayer (or form defects) at the T m, making the membrane leaky to<br />

encapsulated drug.<br />

Therefore, this team arrived at the new tentative liposome design: a 100 nm liposome<br />

comprised of a DPPC bilayer (underlying phase transition at 41.98C) with 4 mol% of<br />

DSPE–PEG (for extended circulation half life) and 10 mol% of the lysolipid MPPC (to enhance<br />

the permeability), containing doxorubicin loaded <strong>by</strong> remote pH loading. The key to the design was<br />

*<br />

SOPC, Stearolyoleoylphosphatidylcholine.<br />

†<br />

MOPC, Monooleoylphosphatidylcholine.<br />

‡<br />

MPPC, Monopalmitoylphosphatidylcholine.<br />

§<br />

DPPC, Dipalmitoylphosphatidylcholine.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Nanoscale Drug Delivery Vehicles for Solid Tumors 697<br />

to work with the recognized anomalous permeability at the Tm (an apparent grain boundary effect in<br />

the melting solid membranes) and to enhance it with a second component (lysolipid). Ultimately, if<br />

this change in composition could change the membrane property of permeability (when heated), the<br />

drug would be freely available, and the performance of the temperature-sensitive liposome in vivo<br />

could be enhanced.<br />

This lysolipid-based design was tested in 100 nm liposomes, and it characterized their ability to<br />

become permeable to small molecules, ions, and the drug doxorubicin. By incorporating 10 mol%<br />

of MPPC or MSPC * into the gel phase bilayers of DPPC liposomes, the permeability rate to<br />

carboxyfluorescein (CF) 148 and dithionite 115 as well as the rate of release of encapsulated drugs<br />

like doxorubicin 115,123,155 was greatly enhanced over that of the pure DPPC bilayer at its phase<br />

transition. For a liposome formulation composed of DPPC:MPPC:DSPE–PEG-2000 † (molar ratio,<br />

90:10:4), 60% of the entrapped content of CF was released in 1 min at 408C, whereas pure DPPC<br />

liposomes released less than 20% CF even in 5 min at 408C. Using the dithionite permeability<br />

assay, 156,157 several lipid systems were compared, providing some very interesting results. In this<br />

assay, the addition of dithionite quenches the fluorescence and produces a decay in the absorbance<br />

spectrum of NBD lipids ‡ that are incorporated in the lipid bilayers at low (1 mol%) concentration<br />

(Figure 34.14). 156 In these experiments, the interaction between dithionite and the NBD-lipid head<br />

groups was recorded as a decay in absorbance at 465 nm in order to avoid attenuation of fluorescence<br />

because of repeated exposure to exciting wavelengths. If the bilayer is permeable, the<br />

absorbance shift will take place on both sides of the bilayer. Otherwise, if the bilayer is impermeable<br />

(e.g., in its gel state), only the outer monolayer is affected. The change in relative absorbance of<br />

the outside monolayer upon addition of dithionite matched its relative lipid distribution (i.e., 54%<br />

outside versus 46% remaining inside), reflecting the high 50 nm radius of curvature of the liposome.<br />

115 As expected, the permeability for pure DPPC was somewhat enhanced at 40, 42, and 438C<br />

in the region of its phase transition compared to lower temperatures of 30 and 378C in the gel phase<br />

region (Figure 34.14). With the inclusion of 10 mol% lysolipid in the bilayers, the absorbance shift<br />

occurred much faster than for pure DPPC alone, showing that the permeability at the phase<br />

transition was dramatically enhanced. 115<br />

Figure 34.15 compares dithionite permeability for several lipid compositions, represented as<br />

permeability rates (per second). The permeability for DPPC at its Tm (41.98C) was enhanced<br />

compared to its gel state, but it was surprisingly only slightly higher than that of a liquid phase<br />

lipid (POPC § ) at 41.98C. 115 At temperatures above 41.98C, both DPPC and POPC exhibited permeability<br />

rates greater that that of DPPC at its phase transition. In contrast, dramatic increases in the<br />

phase transition permeability rate were observed when 10 mol% MPPC or MSPC were incorporated<br />

in the DPPC liposome membranes. The incorporation of these acyl chain-compatible<br />

lysolipids only shifted the bilayer phase transition temperature slightly from 41.98C for DPPC<br />

alone to 41.08C with MPPC and 41.48C with MSPC. 115 Therefore, the incorporation of 10 mol%<br />

lysolipid had achieved the goals of enhancing the transition permeability and slightly lowering the<br />

transition temperature. In line with the expected influence of the acyl chain, the slightly longer<br />

MSPC lipid reduced the Tm less than the MPPC. Differential scanning calorimetry (DSC) studies<br />

showed that the transition region was still very sharp (i.e., not broadened as with the previous<br />

additions of DSPC and Cholesterol). 148<br />

A comparison of the DSC data with the permeability rate data confirms that the grain<br />

boundaries are responsible for the drug release. As shown in Figure 34.16, the permeability rate<br />

rises on the low temperature shoulder of the excess heat flow curve, i.e., before any significant mass<br />

*<br />

MSPC, Monostearoylphosphatidylcholine.<br />

†<br />

DSPE–PEG, Distearoylphosphoethanolamine–Polyethyleneglycol 2000.<br />

‡<br />

Lipids conjugated with the fluorescent molecule NBD, N-(7-nitro-2,1,3-benzoxadiazol).<br />

§ Palmitoyloleoylphosphatidylcholine.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


698<br />

–<br />

SO2 –<br />

SO2 –<br />

SO2 NBD –<br />

NBD –<br />

–<br />

SO2 –<br />

SO2 –NBD<br />

–<br />

SO2 NBD –<br />

–<br />

SO2 – NBD<br />

–<br />

SO2 –NBD<br />

–<br />

SO2 –<br />

SO2 –<br />

SO2 Relative absorbance at 465 nm<br />

1<br />

0.9<br />

0.8<br />

Dithionite added<br />

DPPC 30C<br />

37C<br />

40C<br />

0.7<br />

42C<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

43C<br />

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30<br />

Time (min)<br />

(a) (b) (c)<br />

of solid phase lipid material has melted. Also, the maximum permeability rate coincides with the<br />

midpoint of the transition where it is expected that the grain boundary area is also at its<br />

maximum. 126<br />

As for the specific mechanism for this dramatic enhancement of temperature-triggered permeability<br />

<strong>by</strong> including 10 mol% lysolipid, it was initially hypothesized that desorption of the watersoluble<br />

lysolipid would leave defects in the membrane. However, Mills et al. recently showed that<br />

membrane permeability to dithionite in the transition region was still enhanced following dialysis<br />

against lysolipid-free buffer (with one exchange of buffer) at temperatures above T m for 48 h<br />

(Figure 34.17). 115 Under these conditions, with a massive concentration gradient between liposomes<br />

and dialysate, there was a strong driving force for lysolipid desorption. This study<br />

demonstrated that, contrary to the original hypothesis, lysolipids did not completely desorb from<br />

Relative absorbance at 465 nm<br />

1<br />

0.9<br />

0.8<br />

DPPC:MSPC<br />

Dithionite added<br />

30C<br />

37C<br />

40C<br />

0.7<br />

42C<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

43C<br />

0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30<br />

Time (min)<br />

FIGURE 34.14 The dithionite assay for membrane permeability. (a) Dithionite quenches the fluorescence or,<br />

in this case, shifts the absorbance for NBD lipids incorporated in the bilayer. The outer monolayer is immediately<br />

changed, but the inner monolayer is only affected when the membrane becomes permeable to this ion,<br />

i.e., in the phase transition region. Both compositions show only outer monolayer shifts at temperatures in their<br />

gel state. Pure DPPC (b) shows the anomalous permeability associated with the transition, but for<br />

DPPC:MSPC, (c) the change in absorbance occurs much faster. (From Evans, E. and Needham, D., J. Phys.<br />

Chem., 91 (16), 4219–4228, 1987. With permission.)<br />

Dithionite ion permeability rate (1/min)<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

Nanotechnology for Cancer Therapy<br />

DPPC<br />

DPPC:MPPC (10%)<br />

POPC<br />

DPPC:MSPC (10%)<br />

0<br />

29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52<br />

Temperature ( o C)<br />

FIGURE 34.15 Dithionite ion permeability rates for pure and lysolipid-containing TSL membranes in vitro.<br />

At the phase transition temperature, permeability rates are w10-fold higher for the lysolipid-containing TSLs<br />

(LTSLs) when compared to the pure DPPC bilayer. (From Mills, J. K. and Needham, D., Biochim. Biophys.<br />

Acta, 1716 (2), 77–96, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 699<br />

Heat flow (Endo)<br />

DPPC:MSPC<br />

0.00<br />

30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48<br />

Temperature ( o C)<br />

FIGURE 34.16 Comparison between the DSC thermal profile and the dithionite permeability rate for LTSL<br />

(DPPC:MSPC). This comparison shows that the onset of increased permeability (dashed line) occurs at the low<br />

temperature shoulder of the excess heat curve (solid line). That is, the content release occurs before a large<br />

fraction of the lipid has melted. (From Mills, J. K. and Needham, D., Biochim. Biophys. Acta, 1716 (2), 77–96,<br />

2005. With permission.)<br />

the membrane upon heating to and through the transition; rather, they largely stayed in the<br />

membrane, apparently forming lysolipid-stabilized defects in the bilayer. 115 Because lysolipids<br />

(MPPC and MSPC) have a conical molecular shape as a result of a single acyl chain, they prefer<br />

the micellized state (Figure 34.17b). Therefore, it seems reasonable that they would accumulate at<br />

Dithionite permeability<br />

rate constant (1/min)<br />

(a)<br />

0.4<br />

0.35<br />

0.3<br />

0.25<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

DPPC<br />

DPPC:MPPC10%<br />

DPPC:MPPC (10%)-Following dialysis<br />

0<br />

30 32 34 36 38<br />

Temperature ( o 40 42 44 46 48<br />

C)<br />

shape of lipid<br />

molecule<br />

(b)<br />

Liquid phase region<br />

(c)<br />

0.35<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

packing of lipid<br />

molecules<br />

Permeability rate (1/min) [ ]<br />

water<br />

Interface region<br />

lysolipid-stabilized pore<br />

Gel phase region<br />

lipid<br />

micelle<br />

lipid<br />

bilayer<br />

Gel phase<br />

DPPC<br />

Liquid phase<br />

DPPC<br />

Lysolipid<br />

FIGURE 34.17 (a) Dithionite ion permeability as a function of temperature for freshly prepared LTSLs and<br />

for LTSLs that have been dialysed for 48 h at 458C–488C (liquid state). (b) The maintenance of enhanced<br />

(although slightly reduced) permeability after dialysis shows that some lysolipid must remain in the<br />

membrane. Pure DPPC is shown for comparison. (c) Lysolipids normally exist in aqueous solution as micelles<br />

(left), but they appear to remain present in liposomes at the grain boundaries, possibly stabilizing pores (right).<br />

(From Mills, J. K. and Needham, D., Biochim. Biophys. Acta, 1716 (2), 77–96, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


700<br />

melting grain boundaries, forming semispherical porous edges (Figure 34.17c). The resulting<br />

lysolipid-stabilized pores would form only upon heating to the phase transition region, facilitating<br />

rapid temperature-sensitive release of contents. Previous micropipet studies <strong>by</strong> Zhelev and<br />

Needham on giant lipid vesicles support this hypothesis, suggesting that lysolipid can stabilize<br />

membrane nanopores large enough to pass glucose, but not sucrose, and that this pore formation<br />

does not compromise mechanical stability. 158 Therefore, the incorporation of lysolipids (MPPC or<br />

MSPC) is essential for drug release enhancement for temperature sensitive liposomes.<br />

For the LTSL formulation, doxorubicin (Dox) was readily loaded into preformed gel phase<br />

liposomes <strong>by</strong> remote pH gradient methods (as described in Mayer, Bally, and Cullis and Madden, et<br />

al. but used now at 358C). 115,159,160 The liposomes were then heated, and Dox release was measured<br />

as a function of time and temperature. The same triggered permeability and enhanced release<br />

effects were seen for Dox as with carboxyfluorescein and dithionite. For liposome compositions<br />

of DPPC:MPPC:DSPE–PEG-2000 (90:10:4) and DPPC:MSPC:DSPE–PEG2000 (90:10:4),<br />

w100% and w90%, respectively, of the encapsulated Dox was released within 2 min (at 40 and<br />

41.38C, respectively). 115 This team’s most recent work has improved the experimental conditions in<br />

order to more closely observe drug release in the first few minutes. 160,161 With the incorporation of<br />

8 mol% MPPC in the lipid bilayers, 100% of the drug is released in 5 min whereas with 12 mol%<br />

MPPC, it takes less than 20 s (Figure 34.18).<br />

The all-important temperature dependence for Dox release was also characterized using<br />

10 mol% lysolipid as shown in Figure 34.19. At 30, 37, and 398C, very little drug was released.<br />

At 408C, there was a moderate level of release and at 41.38C, 100% of the drug was released in 4 min.<br />

If this 4 min time point is used to plot the amount released versus temperature (Figure 34.20), 148 then<br />

the optimal temperature range for triggered release is 398C–428C—a temperature window that is<br />

attainable using regional hyperthermia, even for deep-seated tumors such as ovarian and prostate<br />

tumors. 118 This figure also verifies that, as expected, the stealth liposomes do not release drug in a<br />

temperature-triggered fashion. Finally, when the Dox release curve is compared to the DSC curve for<br />

the same liposomes (Figure 34.21), it can be seen that drug release occurs on the low temperature<br />

shoulder of the phase transition, before a significant mass of lipid is melted. As with the dithionite<br />

data shown earlier, this suggests a grain boundary mechanism for the drug permeability.<br />

Importantly, this in vitro data shows that the release time for the lysolipid-containing temperature<br />

sensitive liposome (LTSL) is likely to be faster than the transit time of the liposome through the<br />

tumor circulation. Moreover, release can be triggered at temperatures in the mild clinical<br />

Percent Dox release<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Increasing MSPC<br />

0<br />

0 5 10 15 20 25 30<br />

Time (min)<br />

Nanotechnology for Cancer Therapy<br />

0% MSPC<br />

5% MSPC<br />

7.4% MSPC<br />

8.5% MSPC<br />

9.3% MSPC<br />

9.7% MSPC<br />

11% MSPC<br />

12% MSPC<br />

13% MSPC<br />

15% MSPC<br />

FIGURE 34.18 Drug release as a function of the molar fraction of lysolipid in the bilayer (from 0 mol% to<br />

15 mol%) at 41.58C. Complete content release occurs within 20 s with 12 mol% MSPC. (From Wright, A., In<br />

Mechanical Engineering and Materials Science. Duke University, Durham, NC, <strong>2006</strong>. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 701<br />

4 10 –9<br />

100% Release<br />

3.86 x –9 mol<br />

3.5 10 –9<br />

Moles of DOX released<br />

3 10 –9<br />

2.5 10 –9<br />

2 10 –9<br />

1.5 10 –9<br />

1 10 –9<br />

5 10 –10<br />

0<br />

0<br />

41.3 o C<br />

hyperthermia range. These advantages make the LTSL a much more feasible system for tumor drug<br />

release than any previous liposome. With this in vitro success, this team was very interested to see<br />

what effects the LTSL would have on tumor drug accumulation and tumor growth delay in vivo.<br />

34.3.4 IN VIVO GROWTH DELAY STUDIES<br />

42 o C<br />

45 o C<br />

40 o C<br />

The efficacy of the Dox–LTSL * was first evaluated in a human squamous cell carcinoma (FaDu)<br />

model, showing unprecedented local control for this tumor (in two studies, respectively, showing 11<br />

out of 11 and 6 out of 9 complete regressions, Figure 34.23). 98,123 The FaDu xenografts were grown<br />

in the flanks of nude athymic mice. All animals received an equivalent single dose of 5 mg/kg i.v.<br />

doxorubicin (maximally tolerated dose) and one hour of local hyperthermia at 428C. In addition to<br />

Dox–LTSLCHT, a number of control treatments were tested for comparison. As shown in<br />

Figure 34.22, with saline control, the FaDu tumors took an average of 10 days to reach the end<br />

point of 5 times initial tumor volume, representing a fairly fast growing tumor. 98 Administration of<br />

free Dox showed an average growth time of 12 days, delaying the growth <strong>by</strong> w2 days.Mild<br />

hyperthermia (HT) alone, in the absence of any drug, exhibited a growth time of 18 days. When<br />

free Dox was combined with HT, the growth time was extended <strong>by</strong> another 4 days to an average of<br />

22 days. 98 Therefore, these experiments demonstrate the anti-tumor activity of mild hyperthermia<br />

alone as well as its ability to enhance drug efficacy (as depicted schematically in Figure 34.11).<br />

In the same FaDu studies, doxorubicin was encapsulated in Doxil-like, non-temperature sensitive<br />

liposomes (NTSL) as well as in the lysolipid-containing temperature sensitive liposomes<br />

(LTSL). Figure 34.23a shows the extended growth time of the NTSL in combination with mild<br />

HT (35 days) with all tumors decreasing rapidly in size but eventually growing back. 123,† Compared<br />

39 o C<br />

5 10 15<br />

Time (min)<br />

37 o C<br />

30 o C<br />

20 25 30<br />

FIGURE 34.19 Drug release as a function of temperature. For liposomes with 10 mol% MSPC, the rate of<br />

Dox release increases as a function of temperature with maximum release occurring at the Tm of the lipid<br />

mixture. (From Needham, D. and Dewhirst, M. W., Adv. Drug Deliv.Rev., 53 (3), 285–305, 2001; Needham,<br />

D., et al., Cancer Res., 60 (5), 1197–1201, 2000. With permission.)<br />

*<br />

Dox–LTSL, doxorubicin-encapsulating lysolipid-containing temperature sensitive liposome.<br />

†<br />

The liposome formulations alone, without heat, did not perform much better than the free drug controls, showing that<br />

reduced accumulation and well-entrapped drug is hardly that efficacious.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


702<br />

Percent of DOX released<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

DOXIL<br />

DPPC:MSPC(10%)<br />

30<br />

20<br />

10<br />

0<br />

30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48<br />

Temperature ( o C)<br />

Nanotechnology for Cancer Therapy<br />

Clinical<br />

hyperthermia<br />

range<br />

FIGURE 34.20 Percent of Dox released within 4 min as a function of temperature for DOXIL-like liposomes<br />

and LTSL (DPPC:MSPC 10%). For LTSL, triggered release occurred in the clinically attainable temperature<br />

range of 398C–428C. Over the same temperature range, the Doxil-like liposome (non-temperature-sensitive)<br />

did not release any drug. (From Needham, D. and Dewhirst, M. W., Adv. Drug Deliv. Rev., 53 (3), 285–305,<br />

2001; Needham, D., et al., Cancer Res., 60 (5), 1197–1201, 2000. With permission.)<br />

to NTSL without heat (21 days, not shown), this again proves the synergism of HT and liposomal<br />

drug delivery through increased perfusion and extravasation as demonstrated in previous<br />

studies. 106,109,117,162 The tumor regrowth pattern with NTSL indicates that drug was certainly<br />

delivered, but some cells were able to survive the treatment, perhaps as a result of the relatively<br />

slow leakage of bioavailable drug. The most dramatic result was observed for the LTSL in combination<br />

with HT with complete regressions in all 11 animals for up to 60 days (Figure 34.23b). 123 , *<br />

The importance of rapid drug release was quite apparent. The focus on composition, structure, and<br />

properties for the liposome design and the testing of in vitro performance of the LTSL translated<br />

into a significant improvement in tumor response in vivo.<br />

Following the success of these anti-tumor efficacy studies, additional FaDu experiments quantified<br />

drug delivery to the tumor and to the actual intracellular molecular target. Total drug<br />

deposition, tissue distribution, and fraction bound to DNA/RNA were determined for free Dox<br />

and three different liposome formulations (Doxil-like NTSL † , LTSL, and a more traditional TSL<br />

with a slightly higher melting point). 98 To do this, two extraction procedures were carried out on<br />

tumors harvested just one hour after injection. Extraction with chloroform and silver nitrate identified<br />

all drug in the tumor tissue. Extraction of the tumor tissue with chloroform only identified drug<br />

that was free, protein-bound, or still trapped inside the liposomes. The difference represented the<br />

drug that was bound to the polynucleic acids. Dox concentration was measured using high performance<br />

liquid chromatography (HPLC). The result was that only the LTSL formulation showed<br />

any doxorubicin bound to DNA and RNA (w50% of the total amount in the tumor, as shown in<br />

Figure 34.24). The total tumor drug delivery using LTSL plus HT was 25 ng doxorubicin per mg<br />

tissue (w50 mM). The other liposomes showed 7–8 ng/mg at 428C and w5 ng/mg at 348C. 98 Growth<br />

time directly correlated with the amount of drug deposited. Tumor tissue sections were also visualized<br />

for doxorubicin fluorescence to determine the local distribution of the drug in the tumor and<br />

confirm the relative drug concentrations. 98 The fluorescence intensity was relatively low at 348C<br />

*<br />

The second study again showed superior results with six out of nine complete regressions (one tumor disease free at<br />

40 days) and only two tumors did not respond.<br />

†<br />

NTSL, non-temperature sensitive liposomes.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 703<br />

Heat flow (endo in mW)<br />

0.2<br />

0.15<br />

0.1<br />

0.05<br />

0<br />

0<br />

30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48<br />

Temperature ( o C)<br />

2.5 10 –11<br />

2 10 –11<br />

1.5 10 –11<br />

1 10 –11<br />

5 10 –12<br />

FIGURE 34.21 Comparison between the Dox release rate and the DSC curve for DPPC:MSPC (compare<br />

Figure 34.16). Drug permeability again occurs on the low temperature shoulder of the phase transition before<br />

significant melting of the lipid mass occurs, indicating a grain boundary mechanism for permeability.<br />

after free drug or any liposomal treatment group. At 428C, the LTSL showed extensive fluorescence<br />

(3.09 arbitrary fluorescence units) with the TTSL 3 times less (1.4) and the Doxil-like NTSL 10 times<br />

less (0.31). These fluorescence images show both unencapsulated and encapsulated drug. For the<br />

LTSL, the Dox distribution was pervasive throughout these tumor sections (Figure 34.24, right).<br />

Overall, the LTSL formulation demonstrated increased drug accumulation and triggered drug<br />

bioavailability, leading to complete tumor regressions after a single dose.<br />

Relative tumor volume<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10 days<br />

0 20 40 60<br />

(a) Days<br />

(b)<br />

Relative tumor volume<br />

10<br />

8<br />

6<br />

4<br />

Control 34°C<br />

Control 42°C<br />

Relative tumor volume<br />

Relative tumor volume<br />

2<br />

0<br />

18 days<br />

0 10 20 30 40 50 60<br />

(c) Days<br />

(d)<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

DOX release rate (Moles/s)<br />

Free drug 34°C<br />

12 days<br />

0 10 20 30 40 50 60<br />

Days<br />

Free drug 42°C<br />

22 days<br />

0 20 40 60<br />

Days<br />

FIGURE 34.22 Tumor growth curves (relative tumor volumes) for flank-implanted FaDu xenografts in nude<br />

mice. The treatment groups were (a) control at normothermia (348C); (b) free doxorubicin at 348C (maximally<br />

tolerated dose of 5 mg/kg); (c) 1 h of hyperthermia at 428C; and d) injection of free drug with 1 h HT at 428C.<br />

The endpoint was time to 5 times initial tumor volume or 60 days. Average growth times are shown for each<br />

group. Overall, hyperthermia shows independent anti-tumor activity and enhancement of drug activity. (From<br />

Kong, G., et al., Cancer Res., 60 (24), 6950–6957, 2000. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


704<br />

Relative tumour volume<br />

(V/Vo )<br />

(a)<br />

Relative tumour volume<br />

(V/Vo )<br />

(b)<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

10<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

NLTSL<br />

42°C<br />

35 days<br />

5V o<br />

0 10 20 30<br />

Days<br />

40 50 60<br />

LTSL 42°C<br />

5V o<br />

>60 days<br />

0 10 20 30 40 50 60<br />

Days<br />

FIGURE 34.23 Tumor growth curves (relative tumor volumes) for flank-implanted FaDu xenografts in nude<br />

mice (as in Figure 34.22). All animals received 5 mg/kg doxorubicin and 1 h of local hyperthermia at 428C.<br />

The treatment groups were (a) non-thermal sensitive Doxil-like liposomes plus HT and (b) Dox–LTSLCHT.<br />

The Dox–LTSL showed superior anti-tumor efficacy with 11 of 11 complete regressions (up to 60 days). (From<br />

Needham, D., et al., Cancer Res., 60(5), 1197–1201, 2000. With permission.)<br />

(a)<br />

Dox (ng/mg)<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

–5<br />

NTS-Doxil-like<br />

liposome<br />

LTS liposome<br />

TTS liposome<br />

S34 D34 NT34 TT34 LT34 S42 D42 NT42 TT42 LT42<br />

Treatment group (34°C or 42°C)<br />

Nanotechnology for Cancer Therapy<br />

Amount of Dox bound<br />

to DNA, RNA<br />

Amount of Dox still in<br />

liposomes or bound<br />

to protein<br />

FIGURE 34.24 (a) Dox levels in tumor tissue after i.v. infusion as measured <strong>by</strong> HPLC for each of the<br />

treatment groups. Hatched bars show the amount still encapsulated or bound to protein, and solid bars show<br />

the amount bound to DNA/RNA. Only the Dox–LTSL formulation plus HT (428C) shows significant amounts<br />

of Dox bound to DNA/RNA (w50%). The total tumor [Dox] after LTSL delivery was w50 uM. (right)<br />

Doxorubicin fluorescence in a representative tumor section immediately prepared after treatment with 1 h<br />

HT in combination with (b) NTSL, (c) TTSL, or (d) LTSL. Fluorescent intensities are denoted in arbitrary<br />

units. (From Kong, G., et al., Cancer Res., 60 (24), 6950–6957, 2000. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 705<br />

FIGURE 34.25 A new paradigm for local therapy (compare Figure 34.8). Thermal sensitive liposomes release<br />

drug RAPIDLY (!20 s) when heated to 428C, resulting in drug release in the blood stream. Therefore, LTSL<br />

drug delivery does not require liposome extravasation to deliver free drug to the tumor cell. (From Kong, G.,<br />

et al., Cancer Res., 60 (24), 6950–6957, 2000. With permission.)<br />

Because all three liposomes were 100 nm in size and coated with PEG, they would be expected to<br />

extravasate to the same extent. However, both the total tumor HPLC results and the histologic<br />

sections showed additional accumulation of drug (encapsulated plus free) in the tumor with the<br />

LTSL. This suggested that the greater delivery achieved <strong>by</strong> LTSL may be due to release of the drug in<br />

the blood stream of the heated tumor as schematically depicted in Figure 34.25. The Dox concentration<br />

for even the stealth NTSL (w15 mM) greatly exceeded the 1.8 mMIC50 (media concentration<br />

required to kill 50% of cells in culture) of free Dox measured in the multi-drug resistant human<br />

cervical carcinoma line KB-85. 163 However, the limited NTSL efficacy in these studies together with<br />

the scarcity of complete responses to Doxil in the clinic, suggest that the EPR effect alone has some<br />

major limitations in drug delivery. As previously mentioned, the in vitro drug release time for LTSL<br />

is shorter than the transit time of a liposome through tumor microvasculature; therefore, additional<br />

local delivery through intravascular release would result in free drug diffusion into the tumor tissue.<br />

Furthermore, endothelial cells lining the tumor microvasculature would also be exposed to free<br />

doxorubicin. This pointed to a new mechanism for drug delivery, one that does not completely<br />

rely on the EPR paradigm of liposome extravasation. The possibility of anti-vascular as well as<br />

anti-neoplastic drug action can now be considered as another aspect of this remarkable paradigm shift.<br />

34.4 A NEW PARADIGM FOR DRUG DELIVERY TO TUMOR TISSUES<br />

34.4.1 VASCULAR SHUTDOWN<br />

A series of preclinical studies in window chamber and flank tumors have established that the<br />

mechanism of Dox–LTSL action is actually one of vascular shut down. This is due to the unique<br />

ability of LTSL to intravascularly release Dox as it passes through the blood vessels of the mildly<br />

heated tumor (Figure 34.25). Utilizing the window chamber for in situ red blood cell velocity<br />

measurements, Yuan et al. demonstrated that treatment with LTSL plus HT (1 h at 428C) results in<br />

significantly reduced blood flow immediately after treatment and complete vascular shutdown<br />

within 24 h in all FaDu tumors (Figure 34.26). 54 Before treatment, blood flow velocity was<br />

w0.4 mm/s, and at only 6 h after treatment, blood velocity had slowed to w0.002 mm/s.<br />

Minimal changes in tumor blood flow were seen in all controls, including free drug and empty<br />

LTSL both with and without HT.<br />

As shown in Figure 34.27, direct observation of the blood vessels in the tumors before and after<br />

treatment showed a startling result—not only had blood flow dramatically slowed, but many of the<br />

tumor microvessels had also disappeared. Within 24 h of LTSL plus the 1 h HT treatment, microvessel<br />

density in the tumor was reduced from 3.9 mm/mm 2 to 0.9 mm/mm 2 , whereas the control<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


706<br />

(a)<br />

0.6<br />

LTSL-HT<br />

LTSL-DOX-HT<br />

LTSL-DOX-nHT<br />

LTSL-nHT<br />

(b)<br />

Average RBC velocity<br />

in tumor vessels (mm/s)<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

*<br />

–0.5 0 6 24<br />

Time (h)<br />

* *<br />

treatments again showed little change (Figure 34.27c). 54 Moreover, these effects were only<br />

observed in tumor microvessels; heated normal microvessels were spared. Therefore, this represents<br />

a unique method of functional targeting of therapeutic agents. Measurements of VEGF<br />

and bFGF expression also showed a 50%–60% reduction in growth factor concentrations in treated<br />

tumors. 54 Overall, this study provided very strong evidence for the paradigm shift from EPR to<br />

intravascular release, leading to anti-vascular effects. The EPR effect alone (i.e., extravasation out<br />

of the blood stream) could not achieve this kind of vascular shut down. The unique ability of the<br />

lysolipid-containing bilayer to rapidly transition from a solid wax-like membrane into an extremely<br />

permeable melting state has made possible this rapid intravascular release of drug triggered <strong>by</strong> the<br />

application of only mild hyperthermia. This opens up the possibility for drug delivery patterns that<br />

target both tumor cells and tumor vasculature.<br />

Average RBC velocity<br />

in tumor vessels (mm/s)<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0.0<br />

–0.5 0 6 24<br />

*<br />

Time (h)<br />

LTSL-HT<br />

LTSL-DOX-HT<br />

FIGURE 34.26 Red blood cell (RBC) velocity in FaDu tumor microvessels after injection of DiI-labeled<br />

fluorescent RBCs. (a) For Dox–LTSL plus HT, blood flow was significantly reduced after therapy and was<br />

essentially zero 24 h later. All controls (empty LTSLGHT, free DoxGHT) showed little effect on blood flow<br />

velocity. b) For normal vasculature, Dox–LTSLCHT had a much smaller effect. (From Chen, Q, et al., Mol.<br />

Cancer Ther., With permission.)<br />

(a) FaDu tumor before treatment<br />

(b)<br />

FaDu tumor 24h after treatment<br />

Microvascular density<br />

in tumors (mm/mm 2 )<br />

(c)<br />

Nanotechnology for Cancer Therapy<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

–0.5 0 6<br />

Time (h)<br />

24<br />

*<br />

LTSL-HT<br />

LTSL-DOX-HT<br />

LTSL-DOX-nHT<br />

LTSL-nHT<br />

FIGURE 34.27 Videomicrographs of tumor vasculature showing the tumor blood vessels before (a) and 24 h<br />

after (b) the LTSL plus 1 h HT treatment. After the treatment, almost all blood vessels disappeared with<br />

indications of thrombus formation. (c) A plot of tumor microvessel density versus time after the treatment<br />

comparing LTSLCHT to the same controls as in Figure 34.26. These results point to an anti-vascular<br />

mechanism found only in tumor vessels. Therefore, the Dox–LTSLCHT therapy appears to be anti-vascular<br />

as well as anti-neoplastic. (From Chen, Q., et al., Mol. Cancer Ther., 3 (10), 1311–1317, 2004. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

**


Nanoscale Drug Delivery Vehicles for Solid Tumors 707<br />

34.4.2 EXPLORING OTHER TUMOR MODELS<br />

Dox–LTSL plus mild HT was tested in a number of other tumor models. As with most anti-cancer<br />

therapies, it was found that different tumors responded to LTSL in different ways. Five tumor lines<br />

were treated with saline, HT alone, LTSL alone, or LTSL plus HT. 164 Each group received<br />

equivalent dosing of 6 mg/kg i.v. Dox and/or 1 h HT at 428C. The mouse mammary carcinoma<br />

4T07 was implanted in normal BalbC mice, and four human tumor xenograft lines were implanted<br />

in nude athymic mice: SKOV-3 (ovarian carcinoma), HCT116 (colon carcinoma), FaDu (head and<br />

neck carcinoma), and PC-3 (prostate carcinoma). 164<br />

All control tumors grew rapidly, reaching 5x tumor volume (5xV) in 10–15 days. HT alone had<br />

minimal efficacy in delaying tumor growth, as did LTSL alone without heat. The Dox–LTSL plus<br />

HT groups showed superior tumor growth delay times in all tumor types when compared to all other<br />

treatment groups (p!0.006), but the degree of prolongation was highly variable from one tumor<br />

line to the next (Figure 34.28). For Dox–LTSLCHT, the mean tumor growth times (to 5xV) for<br />

each model were, in order of increasing response, 4T07 (19.8 days)!HCT116 (28.1 days)!FaDu<br />

(41.8 days)!SKOV3 (51.9 days)!PC3(60 days). 164 This trend correlated to the intrinsic growth<br />

rates of each untreated control tumor line, i.e., 4T07 and HCT116 were the fastest growing tumors<br />

and PC was the slowest growing tumor. The fraction of animals remaining disease-free at 60 days<br />

post treatment was also highly variable among tumor lines. The PC3 tumor line showed the greatest<br />

response with 15 out of 20 complete regressions. Of note, the FaDu tumor type showed fewer<br />

complete regressions (1 out of 10) than in previous studies. 98,123<br />

Potential reasons for these variations in LTSL efficacy were explored <strong>by</strong> measuring parameters<br />

describing the tumor microenvironment, tumor cell sensitivity, and overall drug delivery. 164 For<br />

each tumor type, these parameters were compared to the relative growth delay time and the fraction<br />

of animals showing complete regression up to 60 days.<br />

Measurements of pH and pO 2, reflecting the interstitial microenvironment, were performed in<br />

five mice for each tumor line. The hypothesis that pH might be important was based on the fact that<br />

tumors produce excess acid. 165 Because Dox is a weak base and is increasingly less hydrophobic<br />

with decreasing pH, transport of this drug across cell membranes could, in principle, be inhibited <strong>by</strong><br />

low extracellular pH. However, all tumors had the same slightly acidic interstitial pH (6.8–7.0), so<br />

this could not explain the variations among different tumor lines. The pO2 histograms (100–200<br />

measurements per tumor) were obtained using the Oxylite probe as described <strong>by</strong> Braun and<br />

colleagues. 166 It has been reported that hypoxia can reduce the cyotoxicity of Dox and may<br />

contribute to different tumor responses in vivo. 167 In this team’s studies, there were significant<br />

Median Growth Time (Days)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

4T07 HCT116 FaDu SKOV-3 PC-3<br />

FIGURE 34.28 Median growth time (days to five time tumor volume) for each of five tumor types in response<br />

to Dox–LTSLCHT. Ranking of median growth time (MGT) reflects the intrinsic tumor growth rates of these<br />

tumor types. (From Zhao, Y., Yarmolenko, P.S., et al. [In preparation]. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


708<br />

differences in the levels of hypoxia among these tumor lines, ranging from 34 to 64% hypoxic<br />

fraction. It is expected that the least hypoxic (i.e., most oxygenated) tumors would respond better to<br />

therapy. This trend was shown in HCT116, SKOV3, and PC3, but FaDu and 4TO7 deviated from<br />

this pattern (well oxygenated but non-responding). Therefore, the Dox–LTSL efficacy in a given<br />

tumor line was not completely related to its degree of hypoxia. 164<br />

The amount of drug delivered to tumors is known to be an important factor in anti-tumor<br />

efficacy. Therefore, Dox concentration was measured immediately after LTSLCHT treatment in<br />

eight tumors from each tumor line, using previously published HPLC methods. 98 Drug concentration<br />

correlated very well for four out of the five tumors, but HCT116 showed the highest drug<br />

concentration (0.045 mg/mg tumor tissue) along with the poorest response. In addition, cellular<br />

sensitivity to Dox was determined for each tumor line using the clonogenic assay after 24 h of<br />

exposure to LTSLCHT. 164 It was found that the in vivo responses of three tumors correlated well<br />

with in vitro sensitivity to Dox, but two of the most sensitive tumors in vitro (HCT116 and FaDu)<br />

showed poor response. Altogether, there was not a significant relationship between efficacy and<br />

either drug concentration or in vitro drug sensitivity. 164<br />

Overall, the four parameters (extracellular pH, hypoxic fraction, intratumoral drug concentrations,<br />

and in vitro drug sensitivity) that were examined did not yield a clear pattern to explain the<br />

differences in Dox–LTSL efficacy among the tumor types. However, searching for positive correlations,<br />

there were some interesting aspects. PC3 achieved the second highest drug concentration,<br />

aligning with its high number of complete regressions. In addition, PC3 had the slowest intrinsic<br />

(control) growth rate and was the most sensitive tumor line in vitro. On the other hand, HCT116 had<br />

the highest drug concentration, but it yielded only one complete regression (out of 20). The 4T07<br />

tumor yielded the most consistent explanation for its relative lack of efficacy (no regressions); it had<br />

the highest IC10 (least sensitive), lowest drug concentration, and fastest intrinsic growth rate<br />

among all the tumor lines studied. New results from Yuan et al. also show that the relative<br />

amount of LTSL induced vascular damage is much greater in the FaDu line than in the 4T07<br />

line (although 4T07 did show some vascular damage, indicating intravascular release of drug from<br />

liposomes is still an important mechanism). Accordingly, the relative growth delay time (compared<br />

with control) for FaDu was much higher than for 4T07 (FaDu: 41.8 days/12.1 days Z3.5 vs. 4T07:<br />

19.8/7.5Z2.6). Therefore, multiple potential explanations for the relative resistance of 4T07 to this<br />

combined therapy are now available.<br />

These somewhat inconclusive results demonstrate the complexity of liposomal therapeutics for<br />

cancer therapy, and this is expected to extend to any other nanotherapeutic agent. For the Dox–<br />

LTSL, in particular, a new paradigm of drug delivery has now been established, leading to important<br />

questions regarding mechanisms and methods for a given tumor line and even an individual<br />

subject. In addition to the parameters already studied, which describe tumor microenvironment,<br />

drug concentration, and sensitivity of neoplastic cells, it is believed that individual tumor response<br />

may more critically depend on parameters that specifically affect intravascular drug release and<br />

vascular shutdown. It is now imperative to evaluate such parameter, including tumor perfusion,<br />

vascular density, the heatability of vessels, the temperature profile, the location of drug release<br />

(extravascular versus intravascular), the spatial and temporal distribution of the drug, and the<br />

relative drug sensitivity of the tumor vascular endothelium (the new target). A comprehensive<br />

approach is now being taken to evaluate the Dox–LTSL as this is necessary to optimize clinical<br />

therapy. One challenge that has been faced (that, again, extends to all nanotherapeutics) is the lack<br />

of tools for real-time evaluation of liposomal drug delivery and drug release at the scale of the<br />

whole tumor. Though very useful, the window chamber model has been limited to 2D evaluations<br />

and is not applicable to the clinical setting. This team sought to develop a non-invasive method<br />

(such as MRI * ) to image liposomal drug concentration distributions in vivo.<br />

* MRI, magnetic resonance imaging.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Nanoscale Drug Delivery Vehicles for Solid Tumors 709<br />

FIGURE 34.29 Median growth time (days to five times tumor volume) for each of five tumor types in response<br />

to various treatments. The tumor response to Dox–LTSL was greater than all other treatments for all tumor types.<br />

Ranking of median growth time (MGT) is the same regardless of treatment type for these tumors, reflecting<br />

intrinsic tumor growth rates. MGTs bound <strong>by</strong> rectangles are not significantly different from each other<br />

(Wilcoxon). (From Madden, T. D., et al., Chem. Phys. Lipids, 53 (1), 37–46, 1990. With permission.)<br />

34.4.3 REGIONAL DEPOSITION OF DRUGS OBSERVED USING MRI<br />

Through the use of liposomes containing both drug (Dox) and paramagnetic contrast agents<br />

(manganese), Viglianti et al. showed that LTSL content release and distribution within the tumor<br />

can be monitored in vivo using T1-weighted MRI. 168 Furthermore, manganese (Mn) and Dox<br />

concentrations can be calculated using T1 relaxation times and Mn:Dox encapsulation ratios. 168<br />

This method has been validated in a rat fibrosarcoma model, indicating a concordant linear relationship<br />

between T1-based [Dox] and invasive [Dox] as measured <strong>by</strong> HPLC or histologic<br />

fluorescence. 169<br />

The potential of this type of imaging for real-time evaluation of therapeutic protocols is<br />

currently being evaluated, with correlation of drug delivery to outcome on an individual subject<br />

basis. 170 Figure 34.29 shows representative axial MR images from the rat fibrosarcoma model<br />

before (a) and after (b) administration of i.v. Mn/Dox–LTSL plus local HT. The flank tumor is<br />

located in the upper left corner. 169 The catheter at the center of the tumor was used to deliver local<br />

hyperthermia (via heated water circulation). * Liposomes were injected after the tumor had reached<br />

steady state temperatures of 408C–428C. The tumor enhancement pattern shows that this protocol<br />

achieved release of contrast agent (and therefore drug) in the peripheral vasculature,i.e., in the<br />

feeding vessels as soon as liposomes entered the tumor.. This therapeutic schedule also resulted in<br />

rapid accumulation of large amounts of drug, likely due to extensive intravascular release (according<br />

to the new paradigm discussed above). Indeed, the kinetics of drug deposition observed here<br />

were much faster (!10 min to maximum [Dox]) than the kinetics of liposome extravasation with<br />

HT (hours to maximum [Dox] using non-thermally sensitive liposomes). 109 This is not surprising,<br />

since LTSL releases contents within 20 s at 41.38C. 160 Thus, this protocol (inject liposomes during<br />

HT) targets the rapid drug release specifically to the feeding blood vessels of the tumor, which<br />

could maximize the anti-vascular effects. Therefore, we hypothesize that the most effective tumor<br />

response may be achieved with this schedule of therapy.<br />

Current experiments are evaluating different schedules of LTSL plus HT in terms of drug<br />

delivery (spatial and temporal tumor distribution) and tumor response (growth delay time) in the<br />

rat fibrosarcoma model. 170 These results from this kind of study will be important for future clinical<br />

*<br />

This is unlikely to be a clinical protocol, but it is a convenient and MR-compatible method to deliver and control local<br />

heating in animal experiments.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


710<br />

trials, and preliminary indications are that it may be more beneficial to deliver Dox–LTSL in the<br />

hyperthermia facility during steady-state heating. Being able to target not only the tumor but also<br />

the vascular source represents a direct clinical application of this new drug delivery paradigm.<br />

Furthermore, an imageable temperature-sensitive liposome could be applied in the clinic to monitor<br />

content release and so facilitate the control of drug distribution and prediction of tumor response for<br />

a given patient. This new multimodal liposome now provides an anti-neoplastic and anti-vascular<br />

therapy that can be monitored <strong>by</strong> MRI in real time.<br />

34.4.4 CLINICAL TRIALS AND HYPERTHERMIA DEVICES<br />

Based on the in vitro testing and in vivo pre-clinical success described above, clinical trials are now<br />

in progress using the Dox–LTSL formulation in combination with local hyperthermia (ThermoDoxe,<br />

Celsion Corp.). 53 Clinical HT may be delivered to superficial or deep tumors using<br />

microwave, radiofrequency, or ultrasound devices (Figure 34.30). 118 A phase I trial is currently<br />

evaluating Dox–LTSL in combination with radiofrequency thermal ablation (RFA) for liver cancer<br />

(Figure 34.30a). In RFA, the center of the tumor is heated to cytotoxic temperatures greater than<br />

808C. Whereas this treatment clearly kills the center of the tumor, it is found that tumors often recur<br />

at the margins after RFA. This occurs because the temperature profile at the margins is only in the<br />

mild hyperthermia range. Because the Dox–LTSL releases at 398C–428C, it is used in this scenario<br />

primarily to increase cytotoxicity against these boundary cells. In addition, a phase I/II trial will<br />

soon test Dox–LTSL in the setting of chest wall recurrence of breast cancer. This trial will use a<br />

Cool-tip RF<br />

Thermal zone<br />

Surgical margin<br />

Target tissue<br />

(a) (b)<br />

(c) (d)<br />

z (cm)<br />

0<br />

−5<br />

−10<br />

−15<br />

−20<br />

−25<br />

−20<br />

0<br />

20<br />

x (cm)<br />

FIGURE 34.30 (See color insert following page 522.) Example of clinical hyper hyperthermia devices.<br />

(a) The radiofrequency ablation device is percutaneously inserted into the liver tumor under image guidance.<br />

(b) The superficial microwave applicator may be used for cutaneous or subcutaneous tumors (from http://www.<br />

ucsfhyperthermia.org/patientinfo/hyperucsf.htm). (c) The Prolieve transurethral microwave applicator was<br />

developed for benign prostatic hyperplasia but may also be applied for prostate cancer. (d) The RF phased<br />

array device for breast cancer may be integrated into a table so that the patient may be positioned prone above<br />

the device.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

−20<br />

0<br />

20<br />

y (cm)


Nanoscale Drug Delivery Vehicles for Solid Tumors 711<br />

FIGURE 34.31 (See color insert following page 522.) The temperature profile for Celsion’s Prolieve<br />

transurethral microwave applicator shows that this device can deliver 428C at the periphery of the prostate.<br />

This enables the device to be used in combination with Dox–LTSL for prostate cancer that typically occurs in<br />

the periphery.<br />

superficial microwave applicator to apply mild local HT (BSD500e; Figure 34.30b). Other<br />

potential clinical applications include prostate cancer (using the Prolievee transurethral microwave<br />

applicator, Figure 34.30c) and locally advanced breast cancer (using a radiofrequency phased<br />

array, Figure 34.30d). The Prolievee applicator that was originally designed for benign prostatic<br />

hyperplasia has demonstrated an ideal temperature profile for the release of LTSL in the periphery<br />

of the prostate where cancer typically occurs (Figure 34.31). That is, temperatures required for the<br />

release of drug from circulating liposomes can be attained even at the edges of the prostate <strong>by</strong><br />

heating from the urethra.<br />

34.5 CONCLUSIONS<br />

This chapter narrated the development of the lysolipid-containing temperature sensitive liposome<br />

(LTSL) from bench to bedside in an attempt to highlight many of the formulation, clinical, cellular,<br />

and in vivo barriers that must be overcome <strong>by</strong> any nanotechnologically designed cancer therapeutic.<br />

It has shown that this formulation met an unmet need for bioavailable drug <strong>by</strong> combining<br />

previous liposome technologies with a detailed and fundamental knowledge of the materials<br />

science and engineering (i.e., composition, structure, properties, and processing) of lipid vesicle<br />

membranes. Historically, the successes of FDA-approved liposomes included reduced toxicity<br />

compared to free drug administration and a passive tumor accumulation through the EPR effect<br />

(especially for long-circulating Doxil). However, there were still many barriers to effective drug<br />

action, and clinical studies have shown limited improvements in tumor response. Reasoning that<br />

high peaks in bioavailable drug concentration were required for many drugs, this team set out to<br />

design a liposome with rapid triggered release of contents. In the absence of reliable endogenous<br />

triggerable chemistry (perhaps excluding phospholipases), 171 of an external trigger (e.g.,<br />

hyperthermia) was utilized with a focus was on local disease rather than global metastases.<br />

A number of studies contributed to the rationale and design for using hyperthermia to trigger<br />

liposome release were reviewed. In the window chamber tumor model, the synergistic interactions<br />

between hyperthermia and liposomal drug delivery were observed, initially for non-thermal sensitive<br />

liposomes. In addition, many years of fundamental lipid bilayer research revealed important<br />

relationships among liposome composition, nanostructures (such as crystalline defects in the solid<br />

state lipid bilayer), and properties (such as enhanced permeability to drugs at the acyl chain melting<br />

transition). Through the incorporation of lysolipids into the lipid membrane, a dramatic increase in<br />

permeability at the melting phase transition (418C) was achived, resulting in complete release of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


712<br />

encapsulated liposomal contents within 20 seconds was reached. The release mechanism appears to<br />

consist of lysolipid-stabilized pore structures at the interfaces between solid and liquid phase lipids<br />

at the crystalline grain boundaries. This is supported <strong>by</strong> the preservation of liposome permeability<br />

after dialysis and <strong>by</strong> observations of the material structure and properties of giant liposomes.<br />

The preclinical studies that have led to clinical testing of the Dox–LTSL in phase I/II trials were<br />

reviewed. The Dox–LTSL achieved dramatic increases in tumor drug accumulation, up to thirty<br />

times over free Dox, in mice bearing human squamous cell carcinoma (FaDu) xenografts. This<br />

resulted in extensive anti-tumor efficacy (17 out of 20 regressions in two initial studies). Mechanistic<br />

studies in window chamber tumors revealed that Dox–LTSLCHT treatment resulted in<br />

dramatic reduction in tumor blood flow and complete vascular shutdown for tumors tested. This<br />

antivascular action was likely due to intravascular release of Dox from the LTSL, representing a<br />

brand new paradigm for liposomal drug delivery that has previously relied upon extravasation of<br />

the carrier via the EPR effect. Recent studies in other tumor lines have revealed the complexity and<br />

heterogeneity of the tumor response to Dox–LTSLCHT. To monitor liposome release and distribution<br />

in vivo, MRI methods using imageable liposomes have been developed. Ongoing research<br />

will continue to provide insights as this temperature-sensitive nanotechnology moves forward into<br />

clinical trials.<br />

The lessons learned throughout this 15-year effort in triggerable drug delivery are likely to<br />

extend to other nanotechnological therapeutics. Previous studies have shown that it is possible to<br />

retain drug in a submicroscopic carrier or attach it to a nanomolecule. These drug delivery systems<br />

must evade the body’s defenses in order to increase circulation lifetime and tumor uptake.<br />

However, this chapter has categorically shown that it is also essential to release the drug in free<br />

form, either in the blood stream of the tumor or in the perivascular space, so that it is bioavailable to<br />

exert its anti-neoplastic effects as well as important anti-vascular effects.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. National Library of Medicine. ClinicalTrials.gov [cited Feb <strong>2006</strong>]; Available from: http://www.<br />

clinicaltrials.gov/.<br />

2. Gabizon, A. A., Stealth liposomes and tumor targeting: One step further in the quest for the magic<br />

bullet, Clin. Cancer Res., 7(2), 223–225, 2001.<br />

3. Riethmiller, S., From atoxyl to salvarsan: Searching for the magic bullet, Chemotherapy, 51(5),<br />

234–242, 2005.<br />

4. Holland, J. et al., Principles of medical oncology, In Cancer Medicine, Bast, R. et al., Eds., B.C.<br />

Decker Inc., Hamilton, Ont., 2003.<br />

5. Willmott, N. and Daly, J., Microspheres and Regional Cancer Therapy, CRC Press Inc., Boca Raton,<br />

FL, 1994.<br />

6. Almond, B. A. et al., Efficacy of mitoxantrone-loaded albumin microspheres for intratumoral<br />

chemotherapy of breast cancer, J. Control. Release, 91(1–2), 147–155, 2003.<br />

7. Rhines, L. D. et al., Local immunotherapy with interleukin-2 delivered from biodegradable polymer<br />

microspheres combined with interstitial chemotherapy: A novel treatment for experimental malignant<br />

glioma, Neurosurgery, 52(4), 872–879, 2003. discussion 879–880.<br />

8. Allen, T. M. and Cullis, P. R., Drug delivery systems: Entering the mainstream, Science, 303(5665),<br />

1818–1822, 2004.<br />

9. Yuan, F. et al., Microvascular permeability and interstitial penetration of sterically stabilized<br />

(stealth) liposomes in a human tumor xenograft, Cancer Res., 54(13), 3352–3356, 1994.<br />

10. Hobbs, S. K. et al., Regulation of transport pathways in tumor vessels: Role of tumor type and<br />

microenvironment, Proc. Natl Acad. Sci. U.S.A., 95(8), 4607–4612, 1998.<br />

11. Torchilin, V. P., Targeted polymeric micelles for delivery of poorly soluble drugs, Cell. Mol. Life<br />

Sci., 61(19–20), 2549–2559, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 713<br />

12. Wang, J., Mongayt, D., and Torchilin, V. P., Polymeric micelles for delivery of poorly soluble drugs:<br />

Preparation and anticancer activity in vitro of paclitaxel incorporated into mixed micelles based on<br />

poly(ethylene glycol)–lipid conjugate and positively charged lipids, J. Drug Target., 13(1), 73–80,<br />

2005.<br />

13. Kabanov, A. V., Batrakova, E. V., and Alakhov, V. Y., Pluronic block copolymers as novel polymer<br />

therapeutics for drug and gene delivery, J. Control. Release, 82(2–3), 189–212, 2002.<br />

14. Rangel-Yagui, C. O., Pessoa, A. Jr., and Tavares, L. C., Micellar solubilization of drugs, J. Pharm.<br />

Pharm. Sci., 8(2), 147–165, 2005.<br />

15. Allen, T. M., Liposomal drug formulations. Rationale for development and what we can expect for<br />

the future, Drugs, 56(5), 747–756, 1998.<br />

16. Drummond, D. C. et al., Optimizing liposomes for delivery of chemotherapeutic agents to solid<br />

tumors, Pharmacol. Rev., 51(4), 691–743, 1999.<br />

17. Discher, B. M. et al., Polymersomes: Tough vesicles made from diblock copolymers, Science, 284<br />

(5417), 1143–1146, 1999.<br />

18. Dharap, S. S. et al., Molecular targeting of drug delivery systems to ovarian cancer <strong>by</strong> BH3 and<br />

LHRH peptides, J. Control. Release, 91(1–2), 61–73, 2003.<br />

19. Kopecek, J. et al., Water soluble polymers in tumor targeted delivery, J. Control. Release, 74(1–3),<br />

147–158, 2001.<br />

20. Duncan, R. et al., Polymer–drug conjugates, PDEPT and PELT: Basic principles for design and<br />

transfer from the laboratory to clinic, J. Control. Release, 74(1–3), 135–146, 2001.<br />

21. Vasey, P. A. et al., Phase I clinical and pharmacokinetic study of PK1 [N-(2-hydroxypropyl)methacrylamide<br />

copolymer doxorubicin]: First member of a new class of<br />

chemotherapeutic agents–drug–polymer conjugates. Cancer Research Campaign Phase I/II<br />

Committee, Clin. Cancer Res., 5(1), 83–94, 1999.<br />

22. Patri, A. K., Kukowska-Latallo, J. F., and Baker, J. R. Jr., Targeted drug delivery with dendrimers:<br />

Comparison of the release kinetics of covalently conjugated drug and non-covalent drug inclusion<br />

complex, Adv. Drug Deliv. Rev., 57(15), 2203–2214, 2005.<br />

23. Choi, Y. and Baker, J. R. Jr., Targeting cancer cells with DNA-assembled dendrimers: A mix and<br />

match strategy for cancer, Cell Cycle, 4(5), 669–671, 2005.<br />

24. Wong, J. Y. et al., A Phase I trial of 90Y-anti-carcinoembryonic antigen chimeric T84.66 radioimmunotherapy<br />

with 5-fluorouracil in patients with metastatic colorectal cancer, Clin. Cancer Res.,<br />

9(16 Pt 1), 5842–5852, 2003.<br />

25. Juweid, M. E., Radioimmunotherapy of B-cell non-hodgkin’s lymphoma: From clinical trials to<br />

clinical practice, J. Nucl. Med., 43(11), 1507–1529, 2002.<br />

26. Gabizon, A. A. et al., Cardiac safety of pegylated liposomal doxorubicin (Doxil/Caelyx) demonstrated<br />

<strong>by</strong> endomyocardial biopsy in patients with advanced malignancies, Cancer Investig., 22(5),<br />

663–669, 2004.<br />

27. Berry, G. et al., The use of cardiac biopsy to demonstrate reduced cardiotoxicity in AIDS Kaposi’s<br />

sarcoma patients treated with pegylated liposomal doxorubicin, Ann. Oncol., 9(7), 711–716, 1998.<br />

28. Harris, L. et al., Liposome-encapsulated doxorubicin compared with conventional doxorubicin in a<br />

randomized multicenter trial as first-line therapy of metastatic breast carcinoma, Cancer, 94(1),<br />

25–36, 2002.<br />

29. Working, P. K. et al., Reduction of the cardiotoxicity of doxorubicin in rabbits and dogs <strong>by</strong> encapsulation<br />

in long-circulating, pegylated liposomes, J. Pharmacol. Exp. Ther., 289(2), 1128–1133,<br />

1999.<br />

30. Papahadjopoulos, D. et al., Sterically stabilized liposomes: improvements in pharmacokinetics and<br />

antitumor therapeutic efficacy, Proc. Natl Acad. Sci. U.S.A., 88(24), 11460–11464, 1991.<br />

31. Maeda, H. et al., Tumor vascular permeability and the EPR effect in macromolecular therapeutics: A<br />

review, J. Control. Release, 65(1–2), 271–284, 2000.<br />

32. Allen, T. M. et al., Adventures in targeting, J. Liposome Res., 12(1–2), 5–12, 2002.<br />

33. Bendas, G., Immunoliposomes: A promising approach to targeting cancer therapy, BioDrugs, 15(4),<br />

215–224, 2001.<br />

34. Zevalin Injection (Ibritumomab Tiuxetan) Drug Information.[cited; Available from: http://www.<br />

drugs.com/, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


714<br />

Nanotechnology for Cancer Therapy<br />

35. Witzig, T. E. et al., Randomized controlled trial of yttrium-90-labeled ibritumomab tiuxetan radioimmunotherapy<br />

versus rituximab immunotherapy for patients with relapsed or refractory low-grade,<br />

follicular, or transformed B-cell non-Hodgkin’s lymphoma, J. Clin. Oncol., 20(10), 2453–2463,<br />

2002.<br />

36. Witzig, T. E. et al., Safety of yttrium-90 ibritumomab tiuxetan radioimmunotherapy for relapsed<br />

low-grade, follicular, or transformed non-hodgkin’s lymphoma, J. Clin. Oncol., 21(7), 1263–1270,<br />

2003.<br />

37. Nademanee, A. et al., A phase 1/2 trial of high-dose yttrium-90-ibritumomab tiuxetan in combination<br />

with high-dose etoposide and cyclophosphamide followed <strong>by</strong> autologous stem cell<br />

transplantation in patients with poor-risk or relapsed non-Hodgkin lymphoma, Blood, 106 (8),<br />

2896–2902, 2005.<br />

38. Schilder, R. et al., Follow-up results of a phase II study of ibritumomab tiuxetan radioimmunotherapy<br />

in patients with relapsed or refractory low-grade, follicular, or transformed B-cell non-<br />

Hodgkin’s lymphoma and mild thrombocytopenia, Cancer Biother. Radiopharm., 19(4), 478–481,<br />

2004.<br />

39. Wiseman, G. A. et al., Ibritumomab tiuxetan radioimmunotherapy for patients with relapsed or<br />

refractory non-Hodgkin lymphoma and mild thrombocytopenia: A phase II multicenter trial,<br />

Blood, 99(12), 4336–4342, 2002.<br />

40. Gibbs, W. J., Drew, R. H., and Perfect, J. R., Liposomal amphotericin B: Clinical experience and<br />

perspectives, Expert Rev. Anti. Infect. Ther., 3(2), 167–181, 2005.<br />

41. Astellas Pharma, US., Inc. [cited Feb <strong>2006</strong>]; Available from: http://www.astellas.us/press_room/<br />

docs/corp_brochure033105.pdf.<br />

42. Becker, M. J. et al., Enhanced antifungal efficacy in experimental invasive pulmonary aspergillosis<br />

<strong>by</strong> combination of AmBisome with Fungizone as assessed <strong>by</strong> several parameters of antifungal<br />

response, J. Antimicrob. Chemother., 49, 813–820, 2002.<br />

43. Schwartz, R. A., Lambert, W. C., Kaposi Sarcoma, Voorhees, A. V. et al., Eds., eMedicine.com, Inc,<br />

2005.<br />

44. Armstrong, D. K., Relapsed ovarian cancer: Challenges and management strategies for a chronic<br />

disease, Oncologist, 7(suppl. 5), 20–28, 2002.<br />

45. Johnston, S. R. and Gore, M. E., Caelyx: Phase II studies in ovarian cancer, Eur. J. Cancer, 37(suppl.<br />

9), S8–S14, 2001.<br />

46. Gordon, A. N. et al., Recurrent epithelial ovarian carcinoma: A randomized phase III study of<br />

pegylated liposomal doxorubicin versus topotecan, J. Clin. Oncol., 19(14), 3312–3322, 2001.<br />

47. Berek, J. S. and Bast, R. C., Ovarian cancer, In Cancer Medicine, Bast, R. et al., Eds., B.C. Decker<br />

Inc., Hamilton, Ont., 2003.<br />

48. O’Brien, M. E. et al., Reduced cardiotoxicity and comparable efficacy in a phase III trial of pegylated<br />

liposomal doxorubicin HCl (CAELYX/Doxil) versus conventional doxorubicin for first-line treatment<br />

of metastatic breast cancer, Ann. Oncol., 15(3), 440–449, 2004.<br />

49. Muggia, F. M., Clinical efficacy and prospects for use of pegylated liposomal doxorubicin in the<br />

treatment of ovarian and breast cancers, Drugs, 54(suppl. 4), 22–29, 1997.<br />

50. Hauck, M. L., Larue, S. M., et al. Phase I trial of doxorubicin-containing low-temperature sensitive<br />

liposomes in spontaneous canine tumors, Clin. Cancer Res., 12(13) 4004–4010, <strong>2006</strong>.<br />

51. Bangham, A. D., Standish, M. M., and Watkins, J. C., Diffusion of univalent ions across the lamellae<br />

of swollen phospholipids, J. Mol. Biol., 13(1), 238–252, 1965.<br />

52. Gregoriadis, G. et al., Drug-carrier potential of liposomes in cancer chemotherapy, Lancet, 1(7870),<br />

1313–1316, 1974.<br />

53. Celsion Corp. [cited Feb <strong>2006</strong>]; Available from: www.celsion.com/technology/thermodox.cfm.<br />

54. Chen, Q. et al., Targeting tumor microvessels using doxorubicin encapsulated in a novel thermosensitive<br />

liposome, Mol. Cancer Ther., 3(10), 1311–1317, 2004.<br />

55. Allen, T. M. and Smuckler, E. A., Liver pathology accompanying chronic liposome administration in<br />

mouse, Res. Commun. Chem. Pathol. Pharmacol., 50(2), 281–290, 1985.<br />

56. Allen, T. M. et al., Chronic liposome administration in mice: Effects on reticuloendothelial function<br />

and tissue distribution, J. Pharmacol. Exp. Ther., 229(1), 267–275, 1984.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 715<br />

57. Gallois, L., Fiallo, M., and Garnier-Suillerot, A., Comparison of the interaction of doxorubicin,<br />

daunorubicin, idarubicin and idarubicinol with large unilamellar vesicles. Circular dichroism<br />

study, Biochim. Biophys. Acta, 1370(1), 31–40, 1998.<br />

58. Nichols, J. W. and Deamer, D. W., Catecholamine uptake and concentration <strong>by</strong> liposomes maintaining<br />

p/gradients, Biochim. Biophys. Acta, 455(1), 269–271, 1976.<br />

59. Mayer, L. D., Bally, M. B., and Cullis, P. R., Uptake of adriamycin into large unilamellar vesicles in<br />

response to a pH gradient, Biochim. Biophys. Acta, 857(1), 123–126, 1986.<br />

60. Costanzo, L., Board Review Series, Physiology, 2nd ed., Lippincott, Williams and Wilkins, New<br />

York, 1998.<br />

61. Truskey, G. A., Yuan, F., and Katz, D. F., Transport Phenomena in Biological Systems, Pearson<br />

Prentice Hall, Upper Saddle River, NJ, 2004.<br />

62. Wu, N. Z. et al., Measurement of material extravasation in microvascular networks using fluorescence<br />

video-microscopy, Microvasc. Res., 46(2), 231–253, 1993.<br />

63. Takakura, Y., Mahato, R. I., and Hashida, M., Extravasation of macromolecules, Adv. Drug Deliv.<br />

Rev., 34(1), 93–108, 1998.<br />

64. Gerlowski, L. E. and Jain, R. K., Effect of hyperthermia on microvascular permeability to macromolecules<br />

in normal and tumor tissues, Int. J. Microcirc. Clin. Exp., 4(4), 363–372, 1985.<br />

65. Wu, N. Z. et al., Increased microvascular permeability contributes to preferential accumulation of<br />

stealth liposomes in tumor tissue, Cancer Res., 53(16), 3765–3770, 1993.<br />

66. Monsky, W. L. et al., Augmentation of transvascular transport of macromolecules and nanoparticles<br />

in tumors using vascular endothelial growth factor, Cancer Res., 59(16), 4129–4135, 1999.<br />

67. DaunoXome Injection Drug Information. [cited; Available from: http://www.drugs.com/], 2005.<br />

68. Dvorak, H. F. et al., Identification and characterization of the blood vessels of solid tumors that are<br />

leaky to circulating macromolecules, Am. J. Pathol., 133(1), 95–109, 1988.<br />

69. Yuan, F., Transvascular drug delivery in solid tumors, Semin. Radiat. Oncol., 8(3), 164–175, 1998.<br />

70. Yuan, F. et al., Vascular permeability in a human tumor xenograft: Molecular size dependence and<br />

cutoff size, Cancer Res., 55(17), 3752–3756, 1995.<br />

71. Kong, G., Braun, R. D., and Dewhirst, M. W., Hyperthermia enables tumor-specific nanoparticle<br />

delivery: effect of particle size, Cancer Res., 60(16), 4440–4445, 2000.<br />

72. Slater, L. M. et al., Cyclosporin A corrects daunorubicin resistance in ehrlich ascites carcinoma, Br.<br />

J. Cancer, 54(2), 235–238, 1986.<br />

73. Rivera, E., Liposomal anthracyclines in metastatic breast cancer: Clinical update, Oncologist, 8<br />

(Suppl 2), 3–9, 2003.<br />

74. Bradley, A. J. and Devine, D. V., The complement system in liposome clearance: Can complement<br />

deposition be inhibited?, Adv. Drug Deliv. Rev., 32(1–2), 19–29, 1998.<br />

75. Allen, T. M., Hansen, C., and Rutledge, J., Liposomes with prolonged circulation times: Factors<br />

affecting uptake <strong>by</strong> reticuloendothelial and other tissues, Biochim. Biophys. Acta, 981(1), 27–35,<br />

1989.<br />

76. Moghimi, S. M. et al., Non-phagocytic uptake of intravenously injected microspheres in rat spleen:<br />

Influence of particle size and hydrophilic coating, Biochem. Biophys. Res. Commun., 177 (2),<br />

861–866, 1991.<br />

77. Alving, C. R. et al., Therapy of leishmaniasis: Superior efficacies of liposome-encapsulated drugs,<br />

Proc. Natl Acad. Sci. U.S.A., 75(6), 2959–2963, 1978.<br />

78. New, R. R. et al., Antileishmanial activity of antimonials entrapped in liposomes, Nature, 272(5648),<br />

55–56, 1978.<br />

79. Weldon, J. S. et al., Liposomal chemotherapy in visceral leishmaniasis: An ultrastructural study of an<br />

intracellular pathway, Z. Parasitenkd., 69(4), 415–424, 1983.<br />

80. Berman, J. D., U.S. Food and Drug Administration approval of AmBisome (liposomal amphotericin<br />

B) for treatment of visceral leishmaniasis, Clin. Infect. Dis., 28(1), 49–51, 1999.<br />

81. Catania, S. et al., Visceral leishmaniasis treated with liposomal amphotericin B, Pediatr. Infect. Dis.<br />

J., 18(1), 73–74, 1999.<br />

82. Allen, T. M. et al., Liposomes containing synthetic lipid derivatives of poly(ethylene glycol) show<br />

prolonged circulation half-lives in vivo, Biochim. Biophys. Acta, 1066(1), 29–36, 1991.<br />

83. Gabizon, A. and Martin, F., Polyethylene glycol-coated (pegylated) liposomal doxorubicin.<br />

Rationale for use in solid tumours, Drugs, 54(suppl. 4), 15–21, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


716<br />

Nanotechnology for Cancer Therapy<br />

84. Gabizon, A. et al., In vivo fate of folate-targeted polyethylene–glycol liposomes in tumor-bearing<br />

mice, Clin. Cancer Res., 9(17), 6551–6559, 2003.<br />

85. Bedu-Addo, F. K. et al., Effects of polyethyleneglycol chain length and phospholipid acyl chain<br />

composition on the interaction of polyethyleneglycol–phospholipid conjugates with phospholipid:<br />

Implications in liposomal drug delivery, Pharm. Res., 13(5), 710–717, 1996.<br />

86. Torchilin, V. P. et al., Targeted accumulation of polyethylene glycol-coated immunoliposomes in<br />

infarcted rabbit myocardium, FASEB J., 6(9), 2716–2719, 1992.<br />

87. Needham, D., McIntosh, T. J., and Lasic, D. D., Repulsive interactions and mechanical stability of<br />

polymer-grafted lipid membranes, Biochim. Biophys. Acta, 1108(1), 40–48, 1992.<br />

88. Needham, D., McIntosh, T. J., and Zhelev, D. V., Surface chemistry of the sterically stabilized PEGliposome:<br />

General principles, In Liposomes: Rational Design, Janoff, A., Ed., Marcel Dekker, Inc.,<br />

New York, pp. 13–62, 1999.<br />

89. Mori, A. and Huang, L., ”Ninja” liposomes avoiding reticuloendothelial system: Development and<br />

application, In First Shizuoka Drug Delivery System Symposium. Progress in Drug Delivery Systems,<br />

University of Shizuoka, Japan, 1992.<br />

90. Mori, A. et al., Influence of the steric barrier activity of amphipathic poly(ethyleneglycol) and<br />

ganglioside GM1 on the circulation time of liposomes and on the target binding of immunoliposomes<br />

in vivo, FEBS Lett., 284(2), 263–266, 1991.<br />

91. Allen, C. et al., Controlling the physical behavior and biological performance of liposome formulations<br />

through use of surface grafted poly(ethylene glycol), Biosci. Rep., 22(2), 225–250, 2002.<br />

92. Papahadjopoulos, D. and Gabizon, A., Liposomes designed to avoid the reticuloendothelial system,<br />

Prog. Clin. Biol. Res., 343, 85–93, 1990.<br />

93. Vaage, J. et al., Chemoprevention and therapy of mouse mammary carcinomas with doxorubicin<br />

encapsulated in sterically stabilized liposomes, Cancer, 73(9), 2366–2371, 1994.<br />

94. Vaage, J. et al., Therapy of human ovarian carcinoma xenografts using doxorubicin encapsulated in<br />

sterically stabilized liposomes, Cancer, 72(12), 3671–3675, 1993.<br />

95. Vaage, J. et al., Tissue distribution and therapeutic effect of intravenous free or encapsulated<br />

liposomal doxorubicin on human prostate carcinoma xenografts, Cancer, 73(5), 1478–1484, 1994.<br />

96. Vaage, J. et al., Tumour uptake of doxorubicin in polyethylene glycol-coated liposomes and<br />

therapeutic effect against a xenografted human pancreatic carcinoma, Br. J. Cancer, 75 (4),<br />

482–486, 1997.<br />

97. Doxil Injection Drug Information. [cited; Available from: http://www.drugs.com/], 2005.<br />

98. Kong, G. et al., Efficacy of liposomes and hyperthermia in a human tumor xenograft model: Importance<br />

of triggered drug release, Cancer Res., 60(24), 6950–6957, 2000.<br />

99. Needham, D. and Nunn, R. S., Elastic deformation and failure of lipid bilayer membranes containing<br />

cholesterol, Biophys. J., 58(4), 997–1009, 1990.<br />

100. INEX and Enzon announce FDA acceptance of Onco TCS New Drug Application. [cited Feb <strong>2006</strong>];<br />

Available from: http://www.drugs.com/NDA/onco_tcs_040521.html, 2004.<br />

101. Krishna, R. et al., Liposomal and nonliposomal drug pharmacokinetics after administration of liposome-encapsulated<br />

vincristine and their contribution to drug tissue distribution properties,<br />

J. Pharmacol. Exp. Ther., 298(3), 1206–1212, 2001.<br />

102. Gelmon, K. A. et al., Phase I study of liposomal vincristine, J. Clin. Oncol., 17(2), 697–705, 1999.<br />

103. Sarris, A. H. et al., Liposomal vincristine in relapsed non-Hodgkin’s lymphomas: Early results of an<br />

ongoing phase II trial, Ann. Oncol., 11(1), 69–72, 2000.<br />

104. Vincristine liposomal—INEX: Lipid-encapsulated vincristine, onco TCS, transmembrane carrier<br />

system–vincristine, vincacine, vincristine sulfate liposomes for injection, VSLI. Drugs R D, 5(2):<br />

p. 119–23, 2004.<br />

105. Huang, Q. et al., Noninvasive visualization of tumors in rodent dorsal skin window chambers, Nat.<br />

Biotechnol., 17, 1033–1035, 1999.<br />

106. Matteucci, M. L. et al., Hyperthermia increases accumulation of technetium-99m-labeled liposomes<br />

in feline sarcomas, Clin. Cancer Res., 6(9), 3748–3755, 2000.<br />

107. Evans, E. and Needham, D., Physical-properties of surfactant bilayer-membranes–thermal<br />

transitions, elasticity, rigidity, cohesion, and colloidal interactions, J. Phys. Chem., 91(16),<br />

4219–4228, 1987.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 717<br />

108. Kwok, R. and Evans, E., Thermoelasticity of large lecithin bilayer vesicles, Biophys. J., 35(3),<br />

637–652, 1981.<br />

109. Kong, G., Braun, R. D., and Dewhirst, M. W., Characterization of the effect of hyperthermia on<br />

nanoparticle extravasation from tumor vasculature, Cancer Res., 61(7), 3027–3032, 2001.<br />

110. Wu, N. Z. et al., Simultaneous measurement of liposome extravasation and content release in tumors,<br />

Microcirculation, 4(1), 83–101, 1997.<br />

111. Fung, V. W., Chiu, G. N., and Mayer, L. D., Application of purging biotinylated liposomes from<br />

plasma to elucidate influx and efflux processes associated with accumulation of liposomes in solid<br />

tumors, Biochim. Biophys. Acta, 1611(1–2), 63–69, 2003.<br />

112. Satchi-Fainaro, R. et al., PDEPT: Polymer-directed enzyme prodrug therapy. 2. HPMA copolymerbeta-lactamase<br />

and HPMA copolymer-C-Dox as a model combination, Bioconjug. Chem., 14(4),<br />

797–804, 2003.<br />

113. Satchi, R., Connors, T. A., and Duncan, R., PDEPT: Polymer-directed enzyme prodrug therapy. I.<br />

HPMA copolymer-cathepsin B and PK1 as a model combination, Br. J. Cancer, 85(7), 1070–1076,<br />

2001.<br />

114. Needham, D. and Dewhirst, M. W., The development and testing of a new temperature-sensitive<br />

drug delivery system for the treatment of solid tumors, Adv. Drug Deliv. Rev., 53(3), 285–305, 2001.<br />

115. Mills, J. K. and Needham, D., Lysolipid incorporation in dipalmitoylphosphatidylcholine bilayer<br />

membranes enhances the ion permeability and drug release rates at the membrane phase transition,<br />

Biochim. Biophys. Acta, 1716(2), 77–96, 2005.<br />

116. Rensen, P. C. et al., Determination of the upper size limit for uptake and processing of ligands <strong>by</strong> the<br />

asialoglycoprotein receptor on hepatocytes in vitro and in vivo, J. Biol. Chem., 276 (40),<br />

37577–37584, 2001.<br />

117. Kong, G. and Dewhirst, M. W., Hyperthermia and liposomes, Int. J. Hyperthermia, 15(5), 345–370,<br />

1999.<br />

118. Dewhirst, M. et al., Hyperthermia, In Cancer Medicine, Bast, R. et al., Eds., B.C. Decker Inc.,<br />

Hamilton, Ont., 2003.<br />

119. Herman, T. S. et al., Effect of heating on lethality due to hyperthermia and selected chemotherapeutic<br />

drugs, J. Natl Cancer Inst., 68(3), 487–491, 1982.<br />

120. Estimated incidence, mortality, and site of failure of the most common types of cancer in the United<br />

States in 2001. Cancer Facts and Figure 2001 in Atlanta: American Cancer Society. 2001.<br />

121. Suit, H., Assessment of the impact of local control on clinical outcome, Front Radiat. Ther. Oncol.,<br />

29, 17–23, 1996.<br />

122. Schmidt-Ullrich, R., Local tumor control and survival: Clinical evidence and tumor biologic basis,<br />

Surg. Oncol. Clin. N. Am., 9(3), 401–414, 2000. See also page vii.<br />

123. Needham, D. et al., A new temperature-sensitive liposome for use with mild hyperthermia:<br />

Characterization and testing in a human tumor xenograft model, Cancer Res., 60(5), 1197–1201,<br />

2000.<br />

124. Haest, C. W. et al., Fragility of the permeability barrier of Escherichia coli, Biochim. Biophys. Acta,<br />

288(1), 43–53, 1972.<br />

125. Papahadjopoulos, D. et al., Phase transitions in phospholipid vesicles. Fluorescence polarization and<br />

permeability measurements concerning the effect of temperature and cholesterol, Biochim. Biophys.<br />

Acta, 311(3), 330–348, 1973.<br />

126. Mouritsen, O. G., Jorgensen, K., and Honger, T., Permeability of lipid bilayers near the phase<br />

transition, In Permeability and Stability of Lipid Bilayers, Disalvo, E. A. and Simon, S. A., Eds.,<br />

CRC Press, Boca Raton, pp. 137–160, 1995.<br />

127. Mouritsen, O. G. and Zuckermann, M. J., Model of interfacial melting, Phys. Rev. Lett., 58(4),<br />

389–392, 1987.<br />

128. Marsh, D., Watts, A., and Knowles, P. F., Evidence for phase boundary lipid. Permeability of tempocholine<br />

into dimyristoylphosphatidylcholine vesicles at the phase transition, Biochemistry, 15,<br />

3570–3578, 1976.<br />

129. Corvera, E. et al., The permeability and the effect of acyl-chain length for phospholipid bilayers<br />

containing cholesterol: Theory and experiment, Biochim. Biophys. Acta, 1107, 261–270, 1991.<br />

130. Nagle, J. F. and Scott, H. L. Jr., Lateral compressibility of lipid mono- and bilayers. Theory of<br />

membrane permeability, Biochim. Biophys. Acta, 513(2), 236–243, 1978.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


718<br />

Nanotechnology for Cancer Therapy<br />

131. Doniach, S., Thermodynamic fluctuations in phospholipid bilayers, J. Chem. Phys., 68, 4912–4916,<br />

1978.<br />

132. Jacobson, K. and Papahadjopoulos, D., Phase transitions and phase separations in phospholipid<br />

membranes induced <strong>by</strong> changes in temperature, pH, and concentration of bivalent cations, Biochemistry,<br />

14(1), 152–161, 1975.<br />

133. Yatvin, M. B. et al., Design of liposomes for enhanced local release of drugs <strong>by</strong> hyperthermia,<br />

Science, 202(4374), 1290–1293, 1978.<br />

134. Magin, R. L. and Niesman, M. R., Temperature-dependent permeability of large unilamellar liposomes,<br />

Chem. Phys. Lipids, 34(3), 245–256, 1984.<br />

135. Gaber, M. H., Effect of bovine serum on the phase transition temperature of cholesterol-containing<br />

liposomes, J. Microencapsul., 15(2), 207–214, 1998.<br />

136. Ipsen, J. H. and Mouritsen, O. G., Modelling the phase equilibria in two-component membranes of<br />

phospholipids with different acyl-chain lengths, Biochim. Biophys. Acta, 944(2), 121–134, 1988.<br />

137. Marsh, D., CRC Handbook of Lipid Bilayers, CRC Press, Boca Raton, 1990.<br />

138. von Dreele, P. H., Estimation of lateral species separation from phase transitions in nonideal twodimensional<br />

lipid mixtures, Biochemistry, 17(19), 3939–3943, 1978.<br />

139. Bloom, M., Evans, E., and Mouritsen, O. G., Physical properties of the fluid lipid-bilayer component<br />

of cell membranes: A perspective, Q. Rev. Biophys., 24(3), 293–397, 1991.<br />

140. Zhelev, D. V. and Needham, D., Tension-stabilized pores in giant vesicles: Determination of pore<br />

size and pore line tension, Biochim. Biophys. Acta, 1147(1), 89–104, 1993.<br />

141. Olbrich, K., Water permeability and mechanical properties of unsaturated lipid membranes and<br />

sarcolemma vesicles, In Mechanical Engineering and Materials Science. Duke University:<br />

Durham, NC, 1997.<br />

142. Needham, D., Stoicheva, N., and Zhelev, D. V., Exchange of monooleoylphosphatidylcholine as<br />

monomer and micelle with membranes containing poly(ethylene glycol)-lipid, Biophys. J., 73(5),<br />

2615–2629, 1997.<br />

143. Needham, D. and Evans, E., Structure and mechanical properties of giant lipid (DMPC) vesicle<br />

bilayers from 20 degrees C below to 10 degrees C above the liquid crystal-crystalline phase transition<br />

at 24 degrees C, Biochemistry, 27(21), 8261–8269, 1988.<br />

144. Needham, D., McIntosh, T. J., and Evans, E., Thermomechanical and transition properties of dimyristoylphosphatidylcholine/cholesterol<br />

bilayers, Biochemistry, 27(13), 4668–4673, 1988.<br />

145. Kim, D. and Needham, D. Lipid bilayers and monolayers: Characterization using micropipet<br />

manipulation techniques. In Encyclopedia of Surface and Colloid Science, Somasundaran, P.,<br />

Hubbard, A., Et al., Eds., Marcel Dekker, New York, pp. 3057–3086, 2002.<br />

146. Needham, D. and Zhelev, D., The mechanochemistry of lipid vesicles examined <strong>by</strong> micropipet<br />

manipulation techniques, In Vesicles, Rosoff, B., Ed., Macel Dekker, New York, 1996.<br />

147. Needham, D. and Zhelev, D., Use of micropipet manipulation techniques to measure the properties<br />

of giant lipid vesicles, In Giant Vesicles, Luisi, P. L., Walde, P., Eds., Wiley, New York, 2000.<br />

148. Anyarambhatla, G. R. and Needham, D., Enhancement of the phase transition permeability of<br />

DPPC liposomes <strong>by</strong> incorporation of MPPC: A new temperature-sensitive liposome for use with<br />

mild hyperthermia, J. Liposome Res., 9(4), 491–506, 1999.<br />

149. Ickenstein, L. M. et al., Disc formation in cholesterol-free liposomes during phase transition,<br />

Biochim. Biophys. Acta, 1614(2), 135–138, 2003.<br />

150. Kruijff, B. D., Demel, R. A., and Deenen, L. L. M. V., The effect of cholesterol and epicholesterol<br />

incorporation on the permeability and on the phase transition of intact Acholeplasma laidlawii cell<br />

membranes and derived liposomes, Biochim. Biophys. Acta, 255, 331–347, 1972.<br />

151. Kim, D. H. et al., Mechanical properties and microstructure of polycrystalline phospholipid monolayer<br />

shells: Novel solid microparticles, Langmuir, 19(20), 8455–8466, 2003.<br />

152. Simon, S. A. et al., Increased adhesion between neutral lipid bilayers: Interbilayer bridges formed <strong>by</strong><br />

tannic acid, Biophys. J., 66(6), 1943–1958, 1994.<br />

153. Zhelev, D. V. et al., Interaction of synthetic HA2 influenza fusion peptide analog with model<br />

membranes, Biophys. J., 81(1), 285–304, 2001.<br />

154. Zhelev, D. V., Exchange of monooleoylphosphatidylcholine with single egg phosphatidylcholine<br />

vesicle membranes, Biophys. J., 71(1), 257–273, 1996.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoscale Drug Delivery Vehicles for Solid Tumors 719<br />

155. Wright, A. M., Ickenstein, L. M., and Needham, D. Optimization of a thermosensitive liposome<br />

formulation for doxorubicin, in preparation.<br />

156. Langner, M. and Hui, S. W., Dithionite penetration through phospholipid bilayers as a measure of<br />

defects in lipid molecular packing, Chem. Phys. Lipids, 65(1), 23–30, 1993.<br />

157. Langner, M. and Hui, S., Effect of free fatty acids on the permeability of 1,2-dimyristoyl-sn-glycero-<br />

3-phosphocholine bilayer at the main phase transition, Biochim. Biophys. Acta, 1463(2), 439–447,<br />

2000.<br />

158. Zhelev, D. V. and Needham, D., Pore formation and pore dynamics in bilayer membranes, ASME.<br />

Adv. Heat Mass Transfer Biotechnol., 34, 47–49, 1996.<br />

159. Madden, T. D. et al., The accumulation of drugs within large unilamellar vesicles exhibiting a proton<br />

gradient: A survey, Chem. Phys. Lipids, 53(1), 37–46, 1990.<br />

160. Wright, A., Drug loading and release from a thermally sensitive liposome. In Mechanical Engineering<br />

and Materials Science. Duke University, Durham, NC, <strong>2006</strong>.<br />

161. Wright, A. M., Ickenstein, L. M., and Needham, D., Optimization of a thermosensitive liposome<br />

formulation for doxorubicin (in preparation Feb <strong>2006</strong>).<br />

162. Gaber, M. H. et al., Thermosensitive liposomes: Extravasation and release of contents in tumor<br />

microvascular networks, Int. J. Radiat. Oncol. Biol. Phys., 36(5), 1177–1187, 1996.<br />

163. Gaber, M. H., Modulation of doxorubicin resistance in multidrug-resistance cells <strong>by</strong> targeted liposomes<br />

combined with hyperthermia, J. Biochem. Mol. Biol. Biophys., 6(5), 309–314, 2002.<br />

164. Zhao, Y. et al., Comparative effects of a novel thermosensitive doxorubicin-containing liposome in<br />

human and mouse tumors. Mol. Cancer Ther. (to be submitted March <strong>2006</strong>).<br />

165. Owen, C. S., Dependence of proton generation on aerobic or anaerobic metabolism and implications<br />

for tumour pH, Int. J. Hyperthermia, 12(4), 495–499, 1996.<br />

166. Braun, R. D. et al., Comparison of tumor and normal tissue oxygen tension measurements using<br />

OxyLite or microelectrodes in rodents, Am. J. Physiol. Heart Circ. Physiol., 280(6), H2533–H2544,<br />

2001.<br />

167. Teicher, B. A. et al., Classification of antineoplastic treatments <strong>by</strong> their differential toxicity toward<br />

putative oxygenated and hypoxic tumor subpopulations in vivo in the FSaIIC murine fibrosarcoma,<br />

Cancer Res., 50(11), 3339–3344, 1990.<br />

168. Viglianti, B. L. et al., In vivo monitoring of tissue pharmacokinetics of liposome/drug using MRI:<br />

Illustration of targeted delivery, Magn. Reson. Med. 51(6), 1153–62, 2004.<br />

169. Viglianti, B. L. et al., Chemodosimetry of in-vivo tumor liposomal drug concentration using MRI, J.<br />

Magn. Reson. Imaging, in press.<br />

170. Ponce, A. M. et al., Temperature-sensitive liposome release observed <strong>by</strong> MRI: Drug dose painting<br />

with hyperthermia impacts anti-tumor effect. Clin. Cancer Res. in press.<br />

171. Jensen, S. S. et al., Secretory phospholipase A2 as a tumor-specific trigger for targeted delivery of a<br />

novel class of liposomal prodrug anticancer etherlipids, Mol. Cancer Ther., 3(11), 1451–1458, 2004.<br />

172. Liposome. 1999 [cited Feb <strong>2006</strong>]; Available from: http://www.uic.edu/classes/bios/bios100/<br />

lectf03am/liposome.jpg.<br />

173. O’Driscoll, C., and Duncan, R., Chem<strong>by</strong>tes e-zine: Stealth drugs take off 2000 [cited Feb <strong>2006</strong>];<br />

Available from: http://www.chemsoc.org/chem<strong>by</strong>tes/ezine/2000/odriscoll_nov00.htm.<br />

174. Heather Bullen Research <strong>Group</strong>, Northern Kentucky University. [cited Feb <strong>2006</strong>]; Available from:<br />

http://www.nku.edu/wbullenh1/research.htm.<br />

175. Kukowska-Latallo, J. F. et al., Nanoparticle targeting of anticancer drug improves therapeutic<br />

response in animal model of human epithelial cancer, Cancer Res., 65(12), 5317–5324, 2005.<br />

176. Klement, A., TARA News, High-Tech. Behandlung bei malignem Lymphom: Zevalin. 2004 [cited;<br />

Available from: http://www.oeaz.at/zeitung/3aktuell/2004/17/serie/serie17_2004tara.html.<br />

177. Hyperthermia at University of California—San <strong>Francis</strong>co. [cited Feb <strong>2006</strong>.]; Available from: http://<br />

www.ucsfhyperthermia.org/patientinfo/hyperucsf.htm.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


35<br />

CONTENTS<br />

Nanoemulsion Formulations for<br />

Tumor-Targeted Delivery<br />

Sandip B. Tiwari and Mansoor M. Amiji<br />

35.1 Introduction ....................................................................................................................... 723<br />

35.2 Composition of Nanoemulsions ........................................................................................ 725<br />

35.3 Preparation of Nanoemulsions .......................................................................................... 726<br />

35.4 Characterization of Nanoemulsion Systems ..................................................................... 727<br />

35.4.1 Particle Size......................................................................................................... 727<br />

35.4.2 Surface Characteristics ........................................................................................ 728<br />

35.5 Fate of Nanoemulsions In Vivo ........................................................................................ 729<br />

35.5.1 Effect of Particle Size on Disposition and Clearance ........................................ 729<br />

35.5.2 Effect of Surface Characteristics on Disposition and Clearance ....................... 729<br />

35.6 Applications of Nanoemulsions in Cancer Therapy......................................................... 730<br />

35.6.1 Improved Delivery of Lipophilic Drugs ............................................................. 730<br />

35.6.2 Positively Charged Anoemulsions in Cancer Therapy....................................... 731<br />

35.6.3 Photodynamic Therapy of Cancer ...................................................................... 732<br />

35.6.4 Neutron Capture Therapy of Cancer .................................................................. 733<br />

35.6.5 Perfluorochemical Nanoemulsions in Cancer Therapy ...................................... 734<br />

35.7 Conclusions and Future Perspectives................................................................................ 735<br />

References .................................................................................................................................... 736<br />

35.1 INTRODUCTION<br />

An emulsion is generally described as a heterogonous system composed of two immiscible liquids:<br />

one dispersed uniformly as fine droplets throughout the other. The majority of conventional emulsions<br />

in pharmaceutical use have dispersed particles ranging in diameter from 0.1 to 100 mm. As<br />

with other dispersed systems such as suspensions, emulsions are thermodynamically unstable as a<br />

result of the excess free energy associated with the surface of the droplets in the internal phase. The<br />

dispersed droplets strive to come together and reduce the surface area. In addition to this flocculation<br />

effect, the dispersed particles can coalesce, or fuse, and this can result in the eventual<br />

destruction (phase separation) of the emulsion. To minimize this effect, a third component, the<br />

emulsifying agent or surface active agent (surfactant), is added to the system to improve the<br />

stability. The liquid present as small globules is called the dispersed or internal phase, whereas<br />

the liquid where the droplets are dispersed is called the dispersion, continuous, or external phase. In<br />

general, the immiscible liquids are described as oil and water phases. The oil phase may be any<br />

hydrophobic non-polar liquid, and the water phase may be any hydrophilic polar liquid.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

723


724<br />

The term macroemulsion is sometimes employed to distinguish the ordinary emulsions defined<br />

above from microemulsions, nanoemulsions, and micellar systems. Whereas nanoemulsions are<br />

two-phase systems where the dispersed phase droplet size has been made in the nanometer size<br />

range, the microemulsions, and micellar systems are single-phase systems. 1–3 All these systems are a<br />

result of interfacial phenomenon brought out <strong>by</strong> surface active agents. Therefore, it is essential to<br />

understand the molecular mechanisms responsible for the creation of these systems and distinctions<br />

between them. Surface active compounds contain hydrophilic and hydrophobic moieties in the same<br />

molecule. When these surfactants are added to the water phase or the oil phase, one portion of the<br />

molecule is always incompatible with the other solvent molecules. Surfactants overcome this incompatibility<br />

<strong>by</strong> adsorption at interfaces (air–water, oil–water, or solid–water) and/or <strong>by</strong> formation of<br />

aggregates called micelles. Micelles are colloidal aggregates that form above a specific concentration,<br />

the so-called critical micelle concentration (CMC). When the solvent is water, a micellar<br />

aggregate exhibit the structure as shown in Figure 35.1a where the surfactant molecules are oriented<br />

in such a way that the hydrophilic groups are in contact with the water molecules, and the hydrophobic<br />

tails are located in the micellar core away from the aqueous environment. In this case, the core<br />

of a micelle resembles a tiny pool of liquid hydrocarbon where compounds that are poorly soluble in<br />

water but soluble in non-polar solvents can be solubilized. When the solvent is oil phase, inverse (or<br />

reverse) micelles are formed where the surfactant’s hydrophilic groups are located inside<br />

(Figure 35.1b). In this case, the compounds that are soluble in water phase can be solubilized in<br />

the micellar core. The process of solubilization takes place well above the CMC, and because of<br />

solubilization, micelles become swollen and may attain the size of a small droplet (w100 nm).<br />

Consider, for example, that water is solvent, the surfactant concentration is well above CMC, and<br />

the oil phase is also present; the micelles would solubilize large amounts of the oil phase and become<br />

swollen until they start interacting through a phenomenon called percolation. Such packed swollen<br />

micelle structures that could solubilize large amounts of both oil- and water-soluble compounds have<br />

been called microemulsions because they were first thought to be extremely small droplet emulsions.<br />

4 This definition of microemulsions is still controversial with some scientists referring<br />

microemulsions in similar context of emulsions systems (two phase systems, e.g., nanoemulsions)<br />

while others referring to it as a single phase system. 1,4,5 There is, however, general agreement that<br />

microemulsions are optically isotropic, thermodynamically stable systems and are not amenable to<br />

dilution with the external phase as normal emulsions. Application of microemulsions is usually<br />

limited to dermal and peroral application because of their high surfactant concentration that tends<br />

to prove toxic through other routes of administration. Also, they exist in narrow regions of the phase<br />

diagrams and, as such, they are very sensitive to quantitative changes in formulation. A notable<br />

a. Micelle structure<br />

(Dispersion medium is water phase)<br />

Polar head group<br />

Non-polar tail group<br />

Emulsier or surfactant molecule<br />

FIGURE 35.1 (a) Micelles and (b) inverse micelles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

b. Inverse micelle<br />

(Dispersion medium is oil<br />

phase)


Nanoemulsion Formulations for Tumor-Targeted Delivery 725<br />

example of the microemulsion system in pharmaceutical application is Sandimmune Optorale and<br />

Neorale pre-concentrate that contains cyclosporin-A as the active moiety indicated for immune<br />

suppression in organ transplantation. 6<br />

Emulsion formulations can be used orally, parenterally, ophthalmically, or topically, and they<br />

can be used <strong>by</strong> respiratory route. 4,7 Commercial formulations of nanoemulsions are presented in<br />

Table 35.1. Major application areas of nanoemulsions are in the field of total parenteral nutrition<br />

and as carriers for various drugs, anesthetics, and image contrast-enhancement agents. In this<br />

chapter, the discussion on application of nanoemulsion systems is restricted to cancer therapy.<br />

Two types of emulsion systems are possible depending on if the dispersed droplets are oil or water<br />

(i.e., oil-in-water (O/W) or water-in-oil (W/O) emulsion). O/W nanoemulsion systems are of<br />

particular pharmaceutical importance from a parenteral drug delivery point of view. W/O emulsions<br />

are also parenterally used; however, their applications are limited for obtaining the prolonged<br />

release of the water soluble compounds on intramuscular application. Multiple emulsion or double<br />

emulsion systems are also possible where preformed emulsion is re-emulsified with either oil or<br />

water. For example, multiple emulsions of oil-in-water-in-oil (O/W/O) are W/O emulsions where<br />

the water droplets contain dispersed oil globules. Similarly, water-in-oil-in-water (W/O/W)<br />

multiple emulsions are those where the internal and external water phases are separated <strong>by</strong> an<br />

oil phase (Figure 35.2).<br />

35.2 COMPOSITION OF NANOEMULSIONS<br />

Nanoemulsions contain oil phases, surfactants or emulsifiers, active pharmaceutical ingredients<br />

(drugs or diagnostic agents), and additives (Table 35.2). 5,8–10 The oil phases are mainly natural or<br />

synthetic lipids, fatty acids, oils such as medium or long chain triglycerides, or perflurochemicals.<br />

Many oils, in particular, those of vegetable origin, are liable to auto-oxidation, and their use in<br />

pharmaceutical formulations requires the addition of an antioxidant. The most widely used oils for<br />

parenteral applications are purified soybean, corn, castor, peanut, cottonseed, sesame, and safflower<br />

oils. Table 35.1 lists some of the oils that can be used for formulating nanoemulsions. Squalene has<br />

been reported to be the choice of oil for formulating stable nanoemulsions with smallest droplet<br />

size. 11 Squalene, biocompatible oil, is a linear hydrocarbon precursor of cholesterol found in many<br />

tissues, notably the livers of sharks (Squalus) and other fishes. 12 Squalane, a derivative of squalene,<br />

is prepared <strong>by</strong> hydrogenation of squalene and is fully saturated that means that it is not subject to<br />

auto-oxidation, an important issue from a stability point of view. Purified mineral oil is used in<br />

some W/O emulsion preparations. 10 Emulsified perflurochemicals are considered acceptable for<br />

TABLE 35.1<br />

Commercial Nanoemulsion (Sub-Micron Emulsion) Formulations<br />

Drug Brand Manufacturer Therapeutic Indication<br />

Propofol Diprivan Astra Zeneca Anesthetic<br />

Dexamethasone<br />

Palmitate<br />

Limethason Mitsubishi Pharmaceutical,<br />

Japan<br />

Steroid<br />

Alprostadil Liple Mitsubishi Pharmaceutical,<br />

Japan<br />

Vasodilator platelet inhibitor<br />

Flurbiprofen axetil Ropion Kaken Pharmaceuticals,<br />

Japan<br />

Nonsteroidal analgesic<br />

Vitamins A, D, E, K Vitalipid Fresenius Kabi, Europe Parenteral nutrition<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


726<br />

W<br />

Oil<br />

Water-in oil<br />

(W/O emulsion)<br />

Oil<br />

Oil-in water<br />

(O/W emulsion)<br />

intravenous application, and many formulations involving the use of perflurochemicals have<br />

reached the marketplace. 13<br />

The most widely used oils for oral emulsions are various fixed oils of vegetable origin, (corn,<br />

cotton seed, or peanut oil), fish liver oil, and mineral oil. The most commonly used emulsifiers for<br />

parenteral application include either natural lecithins derived from plant or animal sources [e.g.,<br />

egg yolk or soy phosphatidylcholine (PC)] or modified lecithins (e.g., diethylenetriaminepentaacetic<br />

acid (DTPA) conjugated phosphatidylethanolamine (PE) or poly(ethylene glycol) (PEG)modified<br />

lecithin derivatives). The poly(ethylene oxide) (PEO)-containing block copolymers, in<br />

particular Pluronics w or poloxamers, Tetronic w or poloxamines, polysorbates, PEG-conjugated<br />

castor oil derivatives (Cremophore EL w ), polyglycolized as well as glycerides, stearylamine<br />

(SA), 1, 2-dioleoyl-3-trimethylammonium-propane, N-(1, 2-Dioleyl-dihydroxypropyl)-N-(dihydroxy-propyl)<br />

(DOTAP) or other positively charged lipids are also used as emulsifiers or<br />

co-emulsifiers in the nanoemulsion formulations. Other pharmaceutical additives such as tonicity<br />

modifiers, pH adjustment agents, antioxidants, sweeteners, flavors, and preservatives may also be<br />

included in the final formulation, depending on the need.<br />

35.3 PREPARATION OF NANOEMULSIONS<br />

W 2<br />

Water-in-oil-in-water<br />

(W 1 /O/W 2 emulsion)<br />

W<br />

Water<br />

Oil 1<br />

O 2<br />

Oil<br />

Oil-in-water-in-oil<br />

(O 1 /W/O 2 emulsion)<br />

Nanotechnology for Cancer Therapy<br />

Nanoemulsion/ miniemulsion/<br />

sub-micron emulsion<br />

(Type: O/W Nanoemulsion)<br />

As previously discussed, because nanoemulsions are dispersions of fine droplets (in nanometer<br />

range) in a dispersion medium, they are thermodynamically unstable and require considerable<br />

mechanical energy for their preparation. 14–16 The mechanical energy can be supplied in the form<br />

of high pressure homogenizer (e.g., Avestin w homogenizer, Avestin Inc., Ottawa, Canada), Microfluidizer<br />

w (MFIC Corporation, Newton, MA, U.S.A.), or an ultrasonic generator (e.g., VC-505,<br />

Sonics, and Materials Inc., CT, U.S.A.). Stable preparation of nanoemulsions involves the<br />

following steps. First, the active hydrophobic pharmaceutical agent is dissolved in the oil phase<br />

and mixed with the water phase, and the mixture is then emulsified with a high shear homogenizer<br />

for a short period. The coarse oil-in-water emulsion formed is then further homogenized using a<br />

high pressure homogenizer or Microfludizer w . The emulsifier (or surfactant) is either mixed with<br />

the water phase or the oil phase. Incorporation of the surfactant in the oil phase has added advantage<br />

W<br />

Micro emulsion<br />

(Single phase systems)<br />

FIGURE 35.2 The different types of colloidal systems. (Adapted from Salager, J. L., Pharmaceutical Emulsions<br />

and Suspensions, Marcel Dekker, New York, pp. 19–72, 2000.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 727<br />

TABLE 35.2<br />

Components of Nanoemulsion Formulations<br />

as it may aid in enhancing solubility of the lipophillic compound in the oil phase. The size of<br />

nanoemulsion droplets can be manipulated <strong>by</strong> using appropriate pressure settings and number of<br />

cycle runs. Nanoemulsions can also be formulated using direct sonication of the oil phase with<br />

water phase in the presence of emulsifying agent. In this case, the intensity, mode, and duration of<br />

sonication determine the final droplet size. It should, however, be noted that the size of nanoemulsion<br />

droplets is also dependant on the nature of oil (e.g., oil composition and viscosity) and the<br />

emulsifying phase (e.g., type of emulsifier and its concentration).<br />

35.4 CHARACTERIZATION OF NANOEMULSION SYSTEMS<br />

Nanoemulsion systems are routinely characterized for particle size and surface properties (surface<br />

charge and morphology).<br />

35.4.1 PARTICLE SIZE<br />

Oils Emulsifiers<br />

Castor oil Natural lecithins from plant or animal sources<br />

Coconut oil (phospholipids)<br />

Corn oil PEG- phospholipids<br />

Cottonseed oil Poloxamers (e.g. F68)<br />

Evening primrose oil Polysorbates<br />

Fish oil Polyoxyethylene castor oil derivatives<br />

Jojoba oil Polyglycolized glycerides<br />

Lard oil Stearlyamine<br />

Linseed oil Oleylamine<br />

Mineral oil<br />

Olive oil Additives<br />

Peanut oil Antioxidants (a-tocopherol, ascorbic acid)<br />

PEG-vegetable oil Tonicity modifiers (glycerol, sorbitol, xylitol)<br />

Perflurochemicals pH adjustment agents (NaOH or HCl)<br />

Pine nut oil Preservatives<br />

Safflower oil<br />

Sesame oil<br />

Soybean oil<br />

Squalene<br />

Squalane<br />

Sunflower oil<br />

Wheatgerm oil<br />

Source: From Eccleston, G., Encyclopedia of Pharmaceutical Technology, Marcel Dekker, New<br />

York, pp. 137–188, 1992; Chung, H., Kim, T. W., Kwon, M., Kwon, I. C., and Jeong, S. C., J.<br />

Control Release, 71, 339–350, 2001; Allison, A. D., Methods, 19, 87–93, 1999; Klang, S. and<br />

Benita, S., Submicron Emulsions in Drug Targeting and Delivery, Academic Publishers,<br />

Harwood, pp. 119–152, 1998; Lundberg, B. B., J. Pharm. Pharmacol., 49(1), 16–21, 1997;<br />

Campbell, S. E., Kuo, C. J., Hebert, B., Rakusan, K., Marshall, H. W., and Faithfull, N. S.,<br />

Can. J. Cardiol., 7(5), 234–244, 1991.<br />

The size of nanoemulsion droplets determines their behavior both in vitro and in vivo. Particle size<br />

of nanoemulsion droplets can be measured using an ensemble (e.g., spectroscopic methods such as<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


728<br />

light scattering), counting (e.g., microscopy such as freeze fracture electron microscopy) or separation<br />

method (e.g., analytical ultracentrifugation). 17,18 Each method has its own advantages and<br />

disadvantages. For example, when the particle size measurement of standard nanoemulsions (made<br />

with pine nut oil and egg PC) was performed <strong>by</strong> this laboratory using an ensemble and counting<br />

method, the results indicated some discrepancy. The average particle size of the nanoemulsion<br />

droplets was 110G13 nm <strong>by</strong> dynamic light scattering method using ZetaPALS, 90Plus (Brookhaven<br />

Instruments Corporation, Holtsville, NY, U.S.A.). However, the freeze fracture electron<br />

microscopy images indicated the size of the droplets in the range of 25–40 nm with some particle<br />

aggregates in the size range w100–150 nm (Figure 35.3). It is recommended that, in order to obtain<br />

better information on particle size distribution, it is essential that particle size analysis be performed<br />

<strong>by</strong> more than one method. The particle size of the commercial nanoemulsions intended for total<br />

parenteral nutrition is reported to be in the range of 100–400 nm. 19<br />

35.4.2 SURFACE CHARACTERISTICS<br />

Nanotechnology for Cancer Therapy<br />

Similar to particle size, surface charge of the nanoemulsion droplets has marked effect on the<br />

stability of the emulsion system and the droplets’ in vivo disposition and clearance. Conventionally,<br />

the surface charge on the emulsion droplets has been expressed in terms of zeta potential (z). As the<br />

emulsion droplets are a result of interfacial phenomenon brought out <strong>by</strong> surface active agents, their<br />

zeta potential is dependent on the extent of ionization of these surface active agents. According to<br />

DLVO electrostatic theory, the stability of the colloid is a balance between the attractive van der<br />

Waals’ forces and the electrical repulsion because of the net surface charge. 20 If the zeta potential<br />

FIGURE 35.3 Freeze fracture electron microscopy image of paclitaxel nanoemulsions formulated using 20%<br />

oil phase (pine nut oil) and egg phosphatidylcholine as surfactant and deoxycholic acid as co-surfactant. The<br />

inset shows microscopy image at higher magnification. Nanoemulsion droplets were in the size range of<br />

25–40 nm with some particle aggregates in the size range of w100–150 nm. Scale bar in both figures<br />

represents 100 nm.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 729<br />

falls below a certain level, the emulsion droplets will aggregate as a result of the attractive forces.<br />

Conversely, a high zeta potential (either positive or negative), typically more than 30 mV, maintains<br />

a stable system. 8,15,19 The zeta potential of the nanoemulsion droplets is routinely measured<br />

using a Zetasizer (Malvern Instruments, U.K.) or the ZetaPlus instrument (Brookhaven Instruments<br />

Corporation, Holtsville, NY, U.S.A.). The zeta potential of the commercial nanoemulsions used for<br />

total parenteral nutrition is reported to be around K40 to K50 mV at pH 7. 19 The net negative<br />

surface charge is thought to be attributed to the anionic components of the emulsions, mainly the<br />

phospholipids. 21 One can impart positive charge to the nanoemulsions as well using cationic lipids<br />

[e.g., SA, 22–24 oleylamine, 25 2,3-dioleoyloxypropyl-1-trimethylammonium bromide, DOTMA, 26<br />

dimethylaminoethane carbamoyl cholesterol (DC-cholesterol)], 27 polymers (e.g., chitosan), 28,29<br />

and cationic surfactants (e.g., cetyltrimethylammonium bromide). 30<br />

35.5 FATE OF NANOEMULSIONS IN VIVO<br />

Similar to liposomal systems, the oil droplets of a conventional O/W nanoemulsion are rapidly<br />

cleared after intravenous injection <strong>by</strong> the reticuloendothelial system (RES) organs such as the liver,<br />

spleen, and bone marrow. Alternatively, the emulsion droplets would bind to the apolipoproteins<br />

from the blood stream, followed <strong>by</strong> their metabolism similar to chylomicrons.The clearance rate of<br />

emulsion droplets from the systemic circulation is dependant on the physicochemical properties of<br />

the emulsion system (size and surface properties in particular). By appropriate control of these<br />

physicochemical properties, one can manipulate the in vivo behavior of the nanoemulsion systems.<br />

35.5.1 EFFECT OF PARTICLE SIZE ON DISPOSITION AND CLEARANCE<br />

Small-sized emulsion droplets are cleared slower than larger emulsion droplets. It is reported that<br />

small-sized emulsions (100–110 nm) were successful in enhancing the plasma concentration of<br />

rhizoxin although the drug was also distributed to the peripheral tissues. 31 On the other hand,<br />

emulsions with droplet size O200 nm prevented the drug from entering the bone marrow, small<br />

intestine, and other organs not associated with the RES where toxicities of many cytotoxic<br />

compounds are expected. Size of nanoemulsion systems can be controlled <strong>by</strong> processing conditions<br />

(e.g., number of homogenization cycles or duration and intensity of sonication) and also <strong>by</strong> composition<br />

of the emulsifier system. Lundberg has reported that it is possible to control the particle size<br />

of the nanoemulsion system using appropriate co-surfactant, e.g., use of 0.12 wt% of polysorbate<br />

80 in castor oil: egg PC (1:0.4 wt%) emulsion system was found to reduce the size of the droplets<br />

from 50 to 100 nm without using any co-surfactant. 32 The use of Pluronic w F-108 as one of the<br />

co-surfactants also decreased the size of the emulsion droplets. Particle size is also dependant on the<br />

type of the oil phase used for the preparation of the nanoemulsion systems. For example, it was<br />

reported that when squalene was used as the oil phase, smallest nanoemulsion droplets (190 nm)<br />

were observed. 11 Castor, safflower, and lard oils yielded emulsion droplets in the size range of 270–<br />

285 nm, whereas linseed oil yielded droplets with size of around 355 nm.<br />

35.5.2 EFFECT OF SURFACE CHARACTERISTICS ON DISPOSITION AND CLEARANCE<br />

As previously discussed, emulsion droplets are rapidly taken up <strong>by</strong> the cells of the RES. The use of<br />

PEG-modified phospholipids as emulsifies were shown to reduce the uptake <strong>by</strong> the cells of the RES<br />

and to improve the blood circulation times. For example, it was reported that inclusion of PEG<br />

derivatives such as Tween w 80 or PEG–PE into the emulsions composed of castor oil and PC<br />

decreases RES uptake and improves systemic circulation time. 33 Hodoshima et al. 34 reported that<br />

lipid nanoparticles of lipophillic prodrug 4 0 -O-tetrahydropyranyl doxorubicin with synthetic PEG<br />

(1,900)-lipid derivative and PC enhanced the circulation half of the drug to 4.3 h after intravenous<br />

administration in BALB/c mice. A similar approach has been used to improve tumor delivery of<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


730<br />

rhizoxin. 31,35 When lecithin was replaced with hydrophilic poloxamer 338 in the emulsion formulation,<br />

it was possible to avoid the normal disposition of droplets to the liver and spleen. In fact, one<br />

study reported that one could totally eliminate the liver and spleen uptake <strong>by</strong> using a nonionic<br />

co-emulsifier, the block copolymer poloxamine (P 908). 36 Emulsions prepared with Pluronics w as<br />

co-emulsifiers also enhanced the circulation time of the formulation in vivo. The nature of lipid<br />

component also alters the biodistribution profile of emulsion systems. For example, it was reported<br />

that cholesterol-coated emulsions were rapidly removed from circulation compared to those coated<br />

with sphingomyelin. 37 The surface charge of the nanoemulsion droplets also affects the clearance<br />

rate in vivo. Negatively charged nanoemulsions are cleared faster and showed high RES organ<br />

uptake than neutral or positively charged nanoemulsions. 8 On the other hand, positively charged<br />

nanoemulsions showed an initial accumulation in the lungs followed <strong>by</strong> redistribution to the liver<br />

and spleen. 15 This natural fate of positively charged nanoemulsions might be important in developing<br />

a suitable system for delivery of drugs targeted specifically to the lungs in case of transplant<br />

rejection therapy, lung infections, or lung cancer. In brief, the surface modification of nanoemulsion<br />

systems offer an opportunity for prolonging the blood circulation time of the drug, enhancing the<br />

possibility of interaction with target cells of interest. This is, in particular, important for delivery of<br />

anti-cancer drugs as prolonged exposure of the drug is critical in the treatment of multi-drugresistant<br />

tumors.<br />

35.6 APPLICATIONS OF NANOEMULSIONS IN CANCER THERAPY<br />

35.6.1 IMPROVED DELIVERY OF LIPOPHILIC DRUGS<br />

Nanotechnology for Cancer Therapy<br />

The advantages of formulating various lipophilic anti-cancer drugs in submicron O/W emulsion are<br />

obvious. The oil phase of the emulsion systems can act as a solubilizer for the lipophilic compound.<br />

Therefore, solubility of lipophilic drugs can be significantly enhanced in an emulsion system,<br />

leading to smaller administration volumes compared to an aqueous solution. In addition, because<br />

lipophilic drugs are incorporated within the innermost oil phase, they are sequestered from direct<br />

contact with body fluids and tissues. Lipid emulsions can minimize the pain associated with<br />

intravenously administered drugs <strong>by</strong> exposing the tissues to lower concentrations of the drug or<br />

<strong>by</strong> avoiding a tissue irritating vehicle. This has been demonstrated with propofol, 38 diazepam, 39<br />

methohexital, 40 clarithromycin, 41 etomidate, 42 and a novel cytotoxic agent. 43 Also, because these<br />

systems are similar to chylomicrons, they are well-tolerated and have low incidences of side effects<br />

compared to systems based on organic solvents, pH adjustments, and surface active agents (e.g.,<br />

Cremophor) because there is less chance of drug precipitation upon administration. If the drug is<br />

susceptible to hydrolysis or oxidation, it is protected <strong>by</strong> the non-aqueous environment. Furthermore,<br />

incorporation of anti-cancer drugs in submicron emulsions (with droplet size of 50–200 nm)<br />

with long circulation properties are expected to enhance the tumor accumulation of the drug <strong>by</strong><br />

passive targeting through the enhanced permeability and retention effect. 44,45 It possible to enhance<br />

the tumor accumulation of nanoemulsions with appropriate modification of size or surface functionalization<br />

as previously discussed. Oil-in-water submicron emulsions appear to be a viable<br />

alternative for the intravenous administration of various lipophilic cytotoxic drugs. Several<br />

groups of researchers have reported the submicron emulsion formulations of anti-cancer drugs<br />

for improved efficacy and/or reduced toxicity.<br />

Paclitaxel, a highly potent anti-cancer agent initially extracted from the bark of the Pacific yew,<br />

was entrapped in lipid emulsion droplets with triolein as oil core and dipalmitoylphosphatidylcholine<br />

(DPPC) as the principal emulsifier. The emulsion was stabilized with polysorbate 80 and PEGdipalmitoyl<br />

PE. The results of in vitro anti-proliferative activity of paclitaxel against T47D cells<br />

was retained <strong>by</strong> the drug emulsion with IC50 (drug concentration effective in reducing proliferation<br />

<strong>by</strong> 50%) value of 7 nM as compared to 10 and 35 nM for paclitaxel in liposomes and Cremophor w<br />

EL-ethanol (50:50) formulations, respectively. 46 The incorporation of PEG-derivatized<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 731<br />

phospholipid is expected to enhance the in vivo circulation half life of the formulation, there<strong>by</strong><br />

enhancing the exposure of the drug to the targeted tumor mass.<br />

Another study reported the formulation of filter sterilizable emulsion formulation of paclitaxel<br />

using a-tocopherol as the oil phase and a-tocopherylpolyethyleneglycol-1,000 succinate (TGPS)<br />

and poloxamer 407 as emulsifiers. 47 The emulsion droplet size was O200 nm with high levels of<br />

paclitaxel loading (8–10 mg/mL). The formulation exhibited better efficacy and was more tolerable<br />

when studied in B16 melanoma tumor model in mice. In addition, another group of researchers<br />

developed a submicron emulsion (droplet diameter 150 nm) of paclitaxel with oil blend (triglycerols),<br />

egg PC, Tween w 80, and glycerol solution. The formulation showed cytotoxicity against<br />

HeLa cells with an IC 50 at 30 nM. An anti-tumor agent, valinomycin, was formulated in the<br />

emulsion form using the commercially-available Intralipid 10% soybean oil emulsion used in<br />

parenteral nutrition. Evaluation of this formulation in vivo indicated that the emulsion formulation<br />

produced similarly shaped dose-response curve to that of an aqueous suspension, but the emulsion<br />

formulation required a 30-fold lower dose than the suspension to produce therapeutic similar<br />

effects. 7<br />

In formulation of nanoemulsions, selection of the appropriate oil phase is important because<br />

most of the anti-cancer compounds exhibit poor solubility in the oil phase, especially those with<br />

highly lipophilic oils. Kan et al. determined the solubility of paclitaxel in various oils such as<br />

tributyrin, tricaproin, tricaprylin, corn, soyabean, cotton seed, and mineral oil, and they found that<br />

triacylglycerols with short fatty acid chains (tributyrin and tricaproin) were good solvents for<br />

paclitaxel with solubility of more than 9.00 mg/g as compared to other vegetable oils (range<br />

0.14–0.23 mg/gm). 48 Another approach to enhance the oil solubility of the anti-cancer compounds<br />

is the chemical modification or prodrug formation. Prodrugs with increased oil solubility have been<br />

obtained with such anti-cancer drugs as teniposide, etoposide, camptothecin, and paclitaxel,<br />

whereas amphiphilic derivatives have been prepared from fluorodeoxyuridine. 49 Esterification<br />

with long-chain fatty acids (i.e., oleic acid) has also been reported to increase the oil solubility<br />

of many anti-cancer drugs.<br />

Using submicron oil-in-water systems, one can engineer a multifunctional therapeutic system<br />

<strong>by</strong> combining several drugs in the same carrier (i.e., with an oil soluble drug in the core and an<br />

amphiphilic one in the surface monolayer) for more effective treatment of tumors or other diseased<br />

conditions. Emulsion formulations also show promise in cancer chemotherapy as vehicles for<br />

prolonging the drug release after intramuscular and intratumoral injection (W/O systems) and as<br />

a means of enhancing the transport of anti-cancer drugs via the lymphatic system. 10<br />

35.6.2 POSITIVELY CHARGED ANOEMULSIONS IN CANCER THERAPY<br />

Positively charged nanoemulsions systems are expected to interact with negatively charged cell<br />

surfaces more efficiently, and this aspect of the positively charged nanoemulsions has been<br />

explored for possibility of oligonucleotide delivery to cancer cells. 50–53 Because oligonucleotides<br />

molecules display a polyanioinc character and present a large molecular structure, their ability to<br />

cross cell membranes remains very low. Teixeira et al. 50 formulated cationic O/W emulsions of<br />

oligonucleotides (pdT 16) using medium chain triglycerides, PC, poloxamer, and either monocationic<br />

SA lipid PC–SA emulsion or a polycationic lipid, RPR C 18 (PC–RPRC 18 emulsion). They<br />

studied the biodistribution in leukemic P388/ADR cells grown in the peritoneal cavity after intraperitoneal<br />

administration in mice (Table 35.3). The fate of both the pdT and the marker of the oil<br />

phase (cholesteryl oleate) were determined in the fluid (devoid of cancer cells) and in the P388/<br />

ADR cell pellet. The results indicated that pdT in solution remained only in the fluid and did not<br />

associate at all with the tumor cells. However, pdT injected as an emulsion formulation was<br />

detectable in the cancer cells pellet even after 24 h and in very high proportions (up to 18% of<br />

the injected dose). When the area-under-the-curve values of the concentration versus time profiles<br />

for different formulations were compared, it was observed that RPR C 18 formulations favored an<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


732<br />

TABLE 35.3<br />

Area Under the Percent Tissues–Time Curve After Intraperitoneal Administration Into Mice<br />

Bearing P388/ADR Ascites of pdT 16 in Solution or Associated with PC/SA or PC/RPRC 18<br />

Emulsion<br />

AUC0–24 h (% Injected Dose/min)<br />

Preparations Blood Liver Kidneys Lungs Spleen Cell Pellet<br />

Solution 3,870 16,411 4,725 1,231 2,343 3,095<br />

PC/SA 3,393 16,505 4,214 1,254 2,438 13,634<br />

PC/RPRC 2,984 1,167 2,771 850 2,200 22,592<br />

PC/SA, phosphatidylcholine-stearylamine emulsion; PC/RPRC18, Posphatidylcholine-polycationic lipid RPRC18 emulsion;<br />

pdT16, Oligohexadecathymidylate.<br />

Source: Adapted from Teixeira, H., Dubernet, C., Chacun, H., Rabinovich, L., Boutet, V., Deverre, J. R., Benita, S., and<br />

Couvreur, P., J. Control Release, 89(3), 473–482, 2003.<br />

increased association of pdT to the tumor cells compared to SA (Table 35.3). Because the distribution<br />

of the oil marker in the tumor fluid and the P388/ADR cell pellet could not be correlated with<br />

that of pdT, the authors postulated that pdT was probably not taken through the endocytosis of the<br />

oil droplets but <strong>by</strong> positive charges of the emulsion that probably increased membrane permeability<br />

and allowed the pdT molecules to more efficiently enter the cells. In another study, cationic<br />

emulsions composed of 3b [N-(N 0 ,N 0 -DC-Chol)] and dioleoylphosphatidyl ethanolamine, castor<br />

oil, and Tween w 80 efficiently delivered plasmid DNA into various cancer cells with low toxicity.<br />

Furthermore, the use of chitosan as a condensing agent (for DNA) and subsequent complexation<br />

with cationic emulsion composed of DC-chol enhanced the transfection efficiency in vitro<br />

compared to DNA/emulsion complexes with the same formulation with chitosan. 54 In vivo study<br />

in mice showed that with chitosan enhanced emulsion complexes, the GFP mRNA expression was<br />

prolonged in liver and lung until day 6. Goldstein et al. 55 formulated a monoclonal antibody,<br />

AMB8LK-, cationic emulsion for targeted delivery to the tumors that overexpress H-ferritin<br />

(that is recognized <strong>by</strong> AMB8LK). The results of cell culture studies indicated that the coupling<br />

of AMB8LK-Fab’ fragment to the cationic emulsion increased the cells uptake <strong>by</strong> 50% as<br />

compared to non-Fab’ conjugated cationic emulsion. Similar conjugation of B-cell lymphoma<br />

monoclonal antibody, LL2, to the surface of polyethylene glycol-PE-based lipid emulsion globules<br />

has been reported with improved binding to the B-lymphoma cells. 56 The results from these studies<br />

indicate significant potential of cationic emulsion systems as effective drug/gene delivery vehicles<br />

to tumor cells.<br />

35.6.3 PHOTODYNAMIC THERAPY OF CANCER<br />

Nanotechnology for Cancer Therapy<br />

Photodynamic therapy (PDT) of cancer is based on the concept that certain photosensitizers can be<br />

localized in the neoplastic tissue, and subsequently, these photosensitizers can be activated with the<br />

appropriate wavelength (energy) of light to generate active molecular species such as free radicals<br />

and singlet oxygen ( 1 O2) that are toxic to cells and tissues. 57–59 PDT is a binary therapy, and it has a<br />

potential advantage in its inherent dual selectivity. First, selectivity is achieved <strong>by</strong> an increased<br />

concentration of the photosensitizer in target tissue; second, the irradiation can be limited to a<br />

specified volume. Provided that the photosensitizer is nontoxic, only the irradiated areas will be<br />

affected even if the photosensitizer does bind to normal tissues. Selectivity can be further enhanced<br />

<strong>by</strong> binding photosensitizers to molecular delivery systems that have high affinity for target tissue.<br />

Many photosensitizers, however, are characterized <strong>by</strong> a high level of hydrophobicity. The poor<br />

water solubility of most photosensitizers often prevents their direct administration into the<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 733<br />

bloodstream. To overcome this problem, photosensitizers have to be either chemically modified or<br />

incorporaed into lipid delivery systems such as liposomes and emulsions.<br />

Morgan et al. utilized emulsions prepared with the ethoxy castor oil Cremophor w -EL (CR) and<br />

DPPC liposomes for the administration of several hydrophobic purpurins to rats with an implanted<br />

urothelial tumor. 60 PDT studies showed that tin (IV) etiopurpurin (SnET2) at a dose of 1 mg<br />

kg-winduced 100% tumor cure with both delivery vehicles. At lower doses, SnET2 in CR was<br />

more effective than SnET2 in liposomes. 61 Therapeutic doses of SnET2 in CR caused no phototoxicity<br />

to footpad tissue irradiated 24 h after administration although some phototoxicity was<br />

observed with liposomes. A higher PDT efficacy was observed <strong>by</strong> delivering a ketochlorin in<br />

CR than in Tween w 80. The better response of the RIF tumor (murine mammary sarcoma) to<br />

PDT was ascribed to an increased tumor uptake of the ketochlorin, probably caused <strong>by</strong> a longer<br />

persistence of the drug in the circulation. 62 Similar conclusions were drawn <strong>by</strong> comparing the<br />

pharmacokinetics of several germanium (IV)-octabutoxy–phthalocyanines (GePc) incorporated<br />

into DPPC liposomes or in CR emulsion. The Cremophor-delivered GePcs were cleared from<br />

the blood circulation at a much slower rate than the liposome-delivered GePcs. 63 At the same<br />

time, Cremophor induced a slower and reduced uptake of the GePcs in the liver and spleen, and it<br />

greatly enhanced the uptake in the tumor as compared to liposomes. Furthermore, CR induced a<br />

four-fold higher accumulation of the Pcs in the MS-2 fibrosarcoma, yielding in all cases two to three<br />

times higher tumor-to-muscle (peritumoral tissue) and tumor-to-skin drug concentration ratios. 64<br />

Almost similar results have been reported with silicon (IV)-phthalocyanine (SiPc) CR or liposome<br />

formulation in C57B1/6 mice Lewis lung carcinoma. 65<br />

In brief, various PDT therapies have reported two different vehicles for photosensitizers, a<br />

Cremophor oil emulsion and DPPC liposomal vesicles. The reported pharmacokinetic studies<br />

clearly indicate that the former vehicle yields a significantly larger selectivity of tumor targeting,<br />

mainly as a consequence of an enhanced accumulation in the malignant lesion, and no appreciable<br />

differences in the uptake <strong>by</strong> healthy tissues is observed with Cremophor or DPPC liposomes (except<br />

higher uptake <strong>by</strong> brain tissues with CR formulation). The greater accumulation of Cremophordelivered<br />

photosensitizers has been thought to be attributed to the tendency of this vehicle to release<br />

the associated photosensitizing drug to serum low-density lipoproteins (LDLs) in a highly preferential<br />

amount, and DPPC liposomes transfer the drug to all the components of the lipoprotein class<br />

in aliquots that are proportional to the relative concentration of the individual proteins. 61 LDLs are<br />

known to display a preferential interaction with a variety of neoplastic cells through a receptormediated<br />

endocytotic process. The tumor cells often express an increased number of LDL receptors<br />

as compared with normal cells.<br />

35.6.4 NEUTRON CAPTURE THERAPY OF CANCER<br />

Neutron Capture Therapy (NCT) is a binary radiation therapy modality that brings together two<br />

components that when kept separate, have only minor effects on the cells. The first component is a<br />

stable isotope of boron or gadolinium (Gd) that can be concentrated in tumor cells <strong>by</strong> a suitable<br />

delivery vehicle. The second is a beam of low-energy neutrons. Boron or Gd in or adjacent to the<br />

tumor cells disintegrates after capturing a neutron, and the high energy heavy charged particles<br />

produced through this interaction destroy only the cancer cells in close proximity to it, leaving<br />

adjacent normal cells largely unaffected. 66 The success of NCT relies on the targeting of boron and<br />

Gd-based compounds to the tumor mass and to achieve desirable intracellular concentrations of<br />

these agents. At the present time, there are two targets with NCT, namely glioblastoma (malignant<br />

brain tumor) and malignant melanoma.<br />

Lu and co-workers developed and evaluated a very low-density lipoprotein (VLDL), resembling<br />

phospholipid-submicron emulsion as a carrier system for new cholesterol-based boronated<br />

compound, BCH (anti-cancer boron neutron capture therapy compound), for targeted delivery to<br />

cancer cells. 67 Emulsions were formulated <strong>by</strong> sonication method using triolein, cholesterol,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


734<br />

cholesterol oleate as lipids, egg PC, lysophosphatidylcholine as surfactants, and BCH as active<br />

pharmaceutical ingredient. The BCH was solubilized in the inner oily core of the emulsion droplets<br />

(155G5 nm). The formulation was designed based on the fact that LDL and high-density lipoproteins<br />

(HDL) are natural carriers of cholesteryl esters in the body, and certain human and animal<br />

tumor types have been shown to have elevated LDL-receptor activity primarily because the rapidly<br />

dividing cancer cells require higher amounts of cholesterol to build new cell membranes. It is<br />

expected that VLDL-resembling formulation may mimic the VLDL–LDL biological process for<br />

targeted drug delivery to cancer cells. Cell culture data showed sufficient uptake of BCH in rat 9L<br />

glioma cells (O50 mg boron/g cells). The authors proposed that such VLDL-resembling formulation<br />

may serve as a targeted drug delivery system to cancerous cells in vivo.<br />

Miyamoto et al. prepared the nanoemulsions of Gd complexed with DTPA-distearylamide (SA)<br />

and DTPA-distearylester (SE) using soybean oil and hydrogenated PC and various co-surfactants<br />

(Cremophore w -RH60, Myrj w 53, Myrj w 59, or Brij w 700). 68,69 The biodistribution of the Gd was<br />

studied in melanoma-bearing hamsters after intraperitoneal administration of the formulations. The<br />

co-surfactant-containing emulsions were more rapidly and completely cleared from the abdominal<br />

cavity than the plain emulsion without co-surfactants. When the effect of the co-surfactants on the<br />

biodistribution of Gd from Gd–DTPA–SA-containing emulsions in the standard-Gd formulation<br />

were compared, the Cremophore w -RH60 emulsion exhibited the highest Gd accumulation in the<br />

tumor (107 mg Gd/g tumor), possibly resulting from its small particle size (78 nm) and the stable<br />

coat on the particle surfaces with PEO. The Brij w 700-containing emulsion allowed the highest<br />

blood Gd concentration for a prolonged period. However, it exhibited slower Gd accumulation in<br />

the tumor, only reaching an identical level in comparison with the Cremophore w -RH60 emulsion.<br />

Gd–DTPA–SE emulsions exhibited poor tumor accumulation of Gd.<br />

The same group of researchers evaluated the effect of intravenous administration of Gd-nanoemulsions<br />

on biodistribution and tumor accumulation. 70 The biodistribution data revealed that<br />

the intravenous injection had three advantages over the ip injection, namely, a faster and higher<br />

accumulation of Gd and a more extended retention time of Gd. Two intravenous injections of the<br />

standard-Gd-nanoemulsion (1.5 mg Gd/ml) with a 24 h interval doubled the tumor accumulation of<br />

Gd, resulting in 49.7 mg Gd/g wet tumor 12 h after administration. By using a two-fold Gd-enriched<br />

formulation (3.0 mg Gd/ml) in the repeated administration schedule, the accumulation was<br />

doubled again, reaching 101 mg Gd/g wet tumor. This level was comparable to the maximum level<br />

in the single ip injection previously reported. The authors proposed that intravenous injection could<br />

be an alternative to ip injection as an administration route for Gd.<br />

35.6.5 PERFLUOROCHEMICAL NANOEMULSIONS IN CANCER THERAPY<br />

Nanotechnology for Cancer Therapy<br />

Perfluorocarbon emulsions are being clinically evaluated as artificial oxygen carriers to reduce<br />

allogeneic blood transfusions or to improve tissue oxygenation. 13 Perfluorochemicals are chemically<br />

inert synthetic molecules that primarily consist of carbon and fluorine atoms and are clear, colorless<br />

liquids. They have the ability to physically dissolve significant quantities of many gases, including<br />

oxygen and carbon dioxide. Perfluorochemicals are hydrophobic and are not miscible with water.<br />

Perfluorochemicals have to be emulsified for intravenous use. To mimic the natural oxygen carrying<br />

cells (RBCs), the droplet size of perfluorocarbon emulsions is maintained in sub-micron range<br />

(median diameter !0.2 mm). Egg phospholipid has been used as an emulsifier of choice in these<br />

formulations. The examples of the commercial perfluorocarbon emulsions are Oxygente (Alliance<br />

Pharmaceutical Corporation, San Diego, CA, U.S.A.), Oxyfluor w (Hemagen Inc., St Louis, MO,<br />

U.S.A.) and Fluosol-DA (Alpha Therapeutic Corp., Los Angeles, CA, U.S.A.).<br />

The perfluorochemical nanoemulsions (PFCE) have opened interesting opportunities in cancer<br />

therapy. It is suggested that fluorocarbon emulsions might find a role in photodynamic therapy, both<br />

as carriers for sensitizing dyes and also to maintain tissue oxygenation in hypoxic regions of solid<br />

tumors. The high solubility of oxygen in fluorocarbon emulsions maintains solution oxygen<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 735<br />

tension, optimizing photo-oxidative damage. The hydrophobic anti-cancer drugs can be delivered<br />

to the tumor mass <strong>by</strong> dissolving them in a hydrophobic core of the emulsion. Furthermore, PFCE<br />

can be used as an adjuvant to radiation therapy and/or chemotherapy in the treatment of solid<br />

tumors. 71,72 The rationale for the use of PFCE in this therapeutic setting is that solid tumor masses<br />

contain areas of hypoxia that are therapeutically resistant. Since x-rays and many chemotherapeutic<br />

agents require oxygen to be maximally cytotoxic and most normal tissues are well-oxygenated, the<br />

additional oxygen put in circulation <strong>by</strong> the PFCE should not increase the normal tissue toxicities<br />

produced <strong>by</strong> the various therapies. The preclinical studies have shown very positive effects with<br />

single dose and fractionated radiation in several rodent solid tumor models. Many widely used anticancer<br />

drugs, including anti-tumor alkylating agents and doxorubicin, have shown improved<br />

response <strong>by</strong> PFCE coadministration. 73 Also, local application of toxic doses of PFCEs resulted<br />

in the necrosis of cancer cells. This is especially promising in the treatment of cancers of the head<br />

and neck regions that are currently difficult to treat. 74<br />

Lanza and co-workers developed a perfluorocarbon emulsion that can be used for ultrasound<br />

and magnetic resonance imaging (MRI) detection of fibrin and thrombus. 75 Liquid perfluorocarbon<br />

inside the nanoemulsion core makes it stable and acoustically reflective—a benefit for ultrasound<br />

imaging. For MRI, the nanoemulsion can carry Gd molecules. The authors estimates that one<br />

nanoemulsion droplet can carry up to 100,000 Gd molecules, enough for a large MR signal and<br />

effective imaging of the targeted area. The fluorine component allows quantification of drug<br />

amounts to the targeted area using MR spectroscopy. The authors speculated that one could<br />

attach radionuclides such as technetium-99m for nuclear imaging or radio-opaque compounds<br />

for contrast enhancement in computed tomography in the nanoemulsion core. The same group<br />

of researchers have also developed an avb3 integrin-targeted paramagnetic nanoemulsions that can<br />

detect and characterize sparse integrin expression on tumor neovasculature induced <strong>by</strong> nascent<br />

melanoma xenografts. 76 Alphan-beta3 integrin is an excellent marker for angiogenesis and, therefore,<br />

is used as the target to dock the nanoemulsion particles. When the emulsion particles are<br />

injected into the body, the homing ligand causes the droplets to seek out a vb 3 integrin. Detecting<br />

a vb 3 integrin provides a direct and early measure of tumor vascularity. Angiogenesis is associated<br />

with a wide variety of pathologies, and <strong>by</strong> targeting angiogenesis, the nanoemulsions can identify<br />

solid tumors and lymphoid cancers at an early stage. Because angiogenesis also occurs with<br />

atherosclerosis and inflammatory diseases, authors feel that such nanoemulsions can be used for<br />

additional diagnostic applications, including imaging early signs of atherosclerosis, restenosis<br />

following angioplasty, and inflammatory diseases such as rheumatoid arthritis.<br />

35.7 CONCLUSIONS AND FUTURE PERSPECTIVES<br />

Nanoemulsion formulations offer several advantages for the delivery of drugs, biologicals, or<br />

diagnostic agents. Traditionally, nanoemulsions have been used in clinics for more than four<br />

decades as total parenteral nutrition fluids. Several other products for drug delivery applications<br />

such as Diprivan w , Liple w , and Ropion w have also reached the marketplace. Although nanoemulsions<br />

are chiefly seen as vehicles for administering aqueous insoluble drugs, they have more<br />

recently received increasing attention as colloidal carriers for targeted delivery of various anticancer<br />

drugs, photosensitizers, neutron capture therapy agents, or diagnostic agents. Because of<br />

their sub-micron size, they can be easily targeted to the tumor area. Moreover, the possibility of<br />

surface functionalization with a targeting moiety has opened new avenues for targeted delivery of<br />

drugs, genes, photosensitizers, and other molecules to the tumor area. Research with perflurochemical<br />

nanoemulsions has shown promising results for the treatment of cancer in conjugation with<br />

other treatment modalities and targeted delivery to the neovasculature. It is expected that further<br />

research and development work will be carried out in the near future for clinical realization of these<br />

targeted delivery vehicles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


736<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. Salager, J. L., Formulation concepts for the emulsion maker, In Pharmaceutical Emulsions and<br />

Suspensions, Nielloud, F. and Marti-Mestres, G., Eds., Marcel Dekker, New York, pp. 19–72, 2000.<br />

2. Tenjarla, S., Microemulsions: An overview and pharmaceutical applications, Crit. Rev. Ther. Drug<br />

Carrier Syst., 16(5), 461–521, 1999.<br />

3. Keipert, S., Siebenbrodt, I., Luders, F., and Bornschein, M., Microemulsions and their potential<br />

pharmaceutical application, Pharmazie, 44(7), 433–444, 1989.<br />

4. Schoot, H., Colloidal dispersions, In Remington: The Science and Practice of Pharmacy, Gennaro, A.,<br />

Lippincott, R., Ed., 20th ed., Williams & Witkins, Maryland, pp. 288–315, 2000.<br />

5. Sarker, D. K., Engineering of nanoemulsions for drug delivery, Curr. Drug Deliv., 2(4), 297–310,<br />

2005.<br />

6. Jores, K., Lipid nanodispersions as drug carrier systems -a physicochemical characterization, http://<br />

sundoc.bibliothek.uni-halle.de/diss-online/04/04H310/prom.pdf, accessed date, Aug. 16, 2005.<br />

7. Buszello, K. and Muller, B., Emulsions as drug delivery systems, In Pharmaceutical Emulsions and<br />

Suspensions, Nielloud, F. and Marti-Mestres, G., Eds., Marcel Dekker, New York, pp. 191–228, 2000.<br />

8. Yang, S. C. and Benita, S., Enhanced absorption and drug targeting <strong>by</strong> positively charged submicron<br />

emulsions, Drug Dev. Res., 50, 476–486, 2000.<br />

9. Inactive ingredient search for approved drug products, http://www.accessdata.fda.gov/scripts/cder/<br />

iig/index.cfm, date accessed, Nov. 05, 2005.<br />

10. Eccleston, G., Emulsions, In Encyclopedia of Pharmaceutical Technology, Swarbrick, J. and Boylan,<br />

J., Eds., Marcel Dekker, New York, pp. 137–188, 1992.<br />

11. Chung, H., Kim, T. W., Kwon, M., Kwon, I. C., and Jeong, S. C., Oil components modulate physical<br />

characteristics and function of the natural oil emulsions as drug or gene delivery system, J. Control<br />

Release, 71, 339–350, 2001.<br />

12. Allison, A. D., Squelene and squalane emulsions as adjuvants, Methods, 19, 87–93, 1999.<br />

13. Lane, T. A., Perfluorochemical-based artificial oxygen carrying red cell substitutes, Transfus. Sci.,<br />

16(1), 19–31, 1995.<br />

14. Benita, S. and Levy, M. Y., Submicron emulsions as colloidal drug carrieres for intravenous administration:<br />

Comprehensive physicochemical characterization, J. Pharm. Sci., 82, 1069–1079, 1993.<br />

15. Klang, S. and Benita, S., Design and evaluation of submicron emulsions as colloidal drug carriers for<br />

intravenous administration, In Submicron Emulsions in Drug Targeting and Delivery, Benita, S., Ed.,<br />

Academic Publishers, Harwood, pp. 119–152, 1998.<br />

16. Klang, S., Frucht-Pery, J., Hoffman, A., and Benita, S., Physicochemical characterization and acute<br />

toxicity evaluation of a positively charged submicron emulsion vehicle, J. Pharm. Pharmacol., 46,<br />

986–993, 1998.<br />

17. Haskell, R., Nanotechnology for drug delivery, www.banyu-zaidan.or.jp/symp/soyaku/haskell.pdf,<br />

date accessed, Oct. 06, 2005.<br />

18. Haskell, R. J., Characterization of submicron systems via optical methods, J. Pharm. Sci., 87(2),<br />

125–129, 1998.<br />

19. Wabel, C., Influence of lecithin on structure and stability of parenteral fat emulsions, http://www2.<br />

chemie.uni-erlangen.de/services/dissonline/data/dissertation/Christoph_Wabel/html/index.html,<br />

accessed date Oct. 09, 2005.<br />

20. Salager, J. L., Emulsion properties and relate know-how and attain them, In Pharmaceutical Emulsions<br />

and Suspensions, Nielloud, F. and Marti-Mestres, G., Eds., Marcel Dekker, New York, pp. 73–125, 2000.<br />

21. Gaysorn, C., Lyons, R. T., Patel, M. V., and Hem, S. L., Effect of surface charge on the stability of<br />

oil/water emulsions during steam sterilization, J. Pharm. Sci., 88(4), 454–458, 1999.<br />

22. Elbaz, E., Zeevi, A., Klang, S., and Benita, S., Positively charged submicron emulsions- a new type of<br />

colloidal drug carrier, Int. J. Pharm., 96, R1–R6, 1993.<br />

23. Klang, S. H., Siganos, C. S., Benita, S., and Frucht-Pery, J., Evaluation of a positively charged<br />

submicron emulsion of piroxicam on the rabbit corneum healing process following alkali burn,<br />

J. Control Release, 57(1), 19–27, 1999.<br />

24. Klang, S. H., Frucht-Pery, J., Hoffman, A., and Benita, S., Physicochemical characterization and acute<br />

toxicity evaluation of a positively-charged submicron emulsion vehicle, J. Pharm. Pharmacol.,<br />

46(12), 986–993, 1994.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 737<br />

25. Gershanik, T., Benzeno, S., and Benita, S., Interaction of a self-emulsifying lipid drug delivery system<br />

with the everted rat intestinal mucosa as a function of droplet size and surface charge, Pharm. Res.,<br />

15(6), 863–869, 1998.<br />

26. Oku, N., Tokudome, Y., Namba, Y., Saito, N., Endo, M., Hasegawa, Y., Kawai, M., Tsukada, H., and<br />

Okada, S., Effect of serum protein binding on real-time trafficking of liposomes with different charges<br />

analyzed <strong>by</strong> positron emission tomography, Biochim. Biophys. Acta, 1280(1), 149–154, 1996.<br />

27. Gao, X. and Huang, L., A novel cationic liposome reagent for efficient transfection of mammalian<br />

cells, Biochem. Biophys. Res. Commun., 179(1), 280–285, 1991.<br />

28. Calvo, P., Alonso, M. J., Vila-Jato, J. L., and Robinson, J. R., Development of positively charged<br />

colloidal drug carrieres: Chitosan coated polyester nanocapsules and sub-micron emulsions, Colloid<br />

Polm. Sci., 275, 46–53, 1997.<br />

29. Jumaa, M. and Muller, B. W., Physicochemical properties of chitosan-lipid emulsions and their<br />

stability during the autoclaving process, Int. J. Pharm., 183(2), 175–184, 1999.<br />

30. Samama, J. P., Lee, K. M., and Biellmann, J. F., Enzymes and microemulsions: Activity and kinetic<br />

properties of liver alcohol dehydrogenase in ionic water in oil microemulsions, Eur. J. Biochem., 163,<br />

609–617, 1987.<br />

31. Kurihara, A., Shibayama, Y., Mizota, A., Yasuno, A., Ikeda, M., and Hisaoka, M., Pharmacokinetics<br />

of highly lipophilic antitumor agent palmitoyl rhizoxin incorporated in lipid emulsions in rats, Biol.<br />

Pharm. Bull., 19(2), 252–258, 1996.<br />

32. Lundberg, B. B., Preparation of drug carrier emulsions stabilized with phosphatidylcholine surfactant<br />

mixtures, J. Pharm. Sci., 83, 72–75, 1994.<br />

33. Liu, F. and Liu, D., Long-circulating emulsions (oil-in-water) as carriers for lipophilic drugs, Pharm.<br />

Res., 12(7), 1060–1064, 1995.<br />

34. Hodoshima, N., Udagawa, C., Ando, T., Fukuyasu, H., Watanabe, T., and Nakabayashi, S., Lipid<br />

nanoparticles for delivering antitumor drugs, Int. J. Pharm., 146, 81–92, 1997.<br />

35. Kurihara, A., Shibayama, Y., Mizota, A., Yasuno, A., Ikeda, M., Sasagawa, K., Kobayashi, T., and<br />

Hisaoka, M., Enhanced tumor delivery and antitumor activity of palmitoyl rhizoxin using stable lipid<br />

emulsions in mice, Pharm. Res., 13(2), 305–310, 1996.<br />

36. Davis, S. S., Washington, C., West, P., Illum, L., Liversidge, G., Sternson, L., and Kirsh, R., Lipid<br />

emulsions as drug delivery systems, Ann. N.Y. Acad. Sci., 507, 75–88, 1987.<br />

37. Arimoto, I., Matsumoto, C., Tanaka, M., Okuhira, K., Saito, H., and Handa, T., Surface composition<br />

regulates clearance from plasma and triolein lipolysis of lipid emulsions, Lipids, 33(8), 773–779,<br />

1998.<br />

38. Krobbuaban, B., Diregpoke, S., Kumkeaw, S., and Tanomsat, M., Comparison on pain on injection of<br />

a small particle size-lipid emulsion of propofol and standard propofol with or without lidocaine,<br />

J. Med. Assoc. Thai., 88(10), 1401–1405, 2005.<br />

39. Thorn-Alquist, A. M., Parenteral use of diazepam in an emulsion formulation. A clinical study, Acta<br />

Anaesthesiol. Scand., 21(5), 400–404, 1977.<br />

40. Westrin, P., Jonmarker, C., and Werner, O., Dissolving methohexital in a lipid emulsion reduces pain<br />

associated with intravenous injection, Anesthesiology, 76(6), 930–934, 1992.<br />

41. Lovell, M. W., Johnson, H. W., Hui, H. W., Cannon, J. B., Gupta, P. K., and Hsu, C. C., Less-painful<br />

emulsion formulations for intravenous administration of clarithomycin, Int. J. Pharm., 109, 45–57,<br />

1994.<br />

42. Stuttmann, H., Doenicke, A., Kugler, J., and Laub, M., A new formulation of etomidate in lipid<br />

emulsion: Bioavailability and venous irritation, Anaesthesist, 38, 421–423, 1989.<br />

43. Glen, J. B. and Hunter, S. C., Pharmacology of an emulsion formulation of ICI 35 868, Br.<br />

J. Anaesth., 56(6), 617–626, 1984.<br />

44. Jang, S. H., Wientjes, M. G., Lu, D., and Au, J. L., Drug delivery and transport to solid tumors, Pharm.<br />

Res., 20(9), 1337–1350, 2003.<br />

45. Mayhew, E. G., Lasic, D., Babbar, S., and Martin, F. J., Pharmacokinetics and antitumor activity of<br />

epirubicin encapsulated in long-circulating liposomes incorporating a polyethylene glycol-derivatized<br />

phospholipid, Int. J. Cancer, 51(2), 302–309, 1992.<br />

46. Lundberg, B. B., A submicron lipid emulsion coated with amphipathic polyethylene glycol for<br />

parenteral administration of paclitaxel (Taxol), J. Pharm. Pharmacol., 49(1), 16–21, 1997.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


738<br />

Nanotechnology for Cancer Therapy<br />

47. Constantinides, P. P., Lambert, K. J., Tustian, A. K., Schneider, B., Lalji, S., Ma, W., Wentzel, B.,<br />

Kessler, D., Worah, D., and Quay, S. C., Formulation development and antitumor activity of a<br />

filter-sterilizable emulsion of paclitaxel, Pharm. Res., 17(2), 175–182, 2000.<br />

48. Kan, P., Chen, Z. B., Lee, C. J., and Chu, I. M., Development of nonionic surfactant/phospholipid o/w<br />

emulsion as a paclitaxel delivery system, J. Control Release, 58(3), 271–278, 1999.<br />

49. Lundberg, B. B., Sub-micron lipid emulsions as carrieres for anticancer drugs, http://www.cmbl.org.<br />

pl/lundberg.pdf, accessed date, Sept. 25, 2005.<br />

50. Teixeira, H., Dubernet, C., Chacun, H., Rabinovich, L., Boutet, V., Deverre, J. R., Benita, S., and<br />

Couvreur, P., Cationic emulsions improves the delivery of oligonucleotides to leukemic P388/ADR<br />

cells in ascite, J. Control Release, 89(3), 473–482, 2003.<br />

51. Teixeira, H., Dubernet, C., Puisieux, F., Benita, S., and Couvreur, P., Submicron cationic emulsions as<br />

a new delivery system for oligonucleotides, Pharm. Res., 16(1), 30–36, 1999.<br />

52. Teixeira, H., Dubernet, C., Rosilio, V., Laigle, A., Deverre, J. R., Scherman, D., Benita, S., and<br />

Couvreur, P., Factors influencing the oligonucleotides release from O-W submicron cationic emulsions,<br />

J. Control Release, 70(1–2), 243–255, 2001.<br />

53. Teixeira, H., Rosilio, V., Laigle, A., Lepault, J., Erk, I., Scherman, D., Benita, S., Couvreur, P., and<br />

Dubernet, C., Characterization of oligonucleotide/lipid interactions in submicron cationic emulsions:<br />

Influence of the cationic lipid structure and the presence of PEG-lipids, Biophys. Chem., 92(3),<br />

169–181, 2001.<br />

54. Lee, M. K., Chun, S. K., Choi, W. J., Kim, J. K., Choi, S. H., Kim, A., Oungbho, K., Park, J. S., Ahn,<br />

W. S., and Kim, C. K., The use of chitosan as a condensing agent to enhance emulsion-mediated gene<br />

transfer, Biomaterials, 26(14), 2147–2156, 2005.<br />

55. Goldstein, D., Nassar, T., Lambert, G., Kadouche, J., and Benita, S., The design and evaluation of a<br />

novel targeted drug delivery system using cationic emulsion-antibody conjugates, J. Control Release,<br />

108(2–3), 418–432, 2005.<br />

56. Lundberg, B. B., Griffiths, G., and Hansen, H., Conjugation of an anti-B-cell lymphoma monoclonal<br />

antibody, LL2, to long-circulating drug-carrier lipid emulsions, J. Pharm. Pharmacol., 51(10),<br />

1099–1105, 1999.<br />

57. Waldow, S. M., Henderson, B. W., and Dougherty, T. J., Hyperthermic potentiation of photodynamic<br />

therapy employing Photofrin I and II: Comparison of results using three animal tumor models, Lasers<br />

Surg. Med., 7(1), 12–22, 1987.<br />

58. Dougherty, T. J., Photosensitizers: therapy and detection of malignant tumors, Photochem. Photobiol.,<br />

45(6), 879–889, 1987.<br />

59. Potter, W. R., Mang, T. S., and Dougherty, T. J., The theory of photodynamic therapy dosimetry:<br />

Consequences of photo-destruction of sensitizer, Photochem. Photobiol., 46(1), 97–101, 1987.<br />

60. Morgan, A. R., Garbo, G. M., Keck, R. W., and Selman, S. H., New photosensitizers for photodynamic<br />

therapy: Combined effect of metallopurpurin derivatives and light on transplantable bladder<br />

tumors, Cancer Res., 48(1), 194–198, 1988.<br />

61. Reddi, E., Role of delivery vehicles for photosensitizers in the photodynamic therapy of tumours,<br />

J. Photochem. Photobiol. B, 37(3), 189–195, 1997.<br />

62. Woodburn, K., Chang, C. K., Lee, S., Henderson, B., and Kessel, D., Biodistribution and PDT efficacy<br />

of a ketochlorin photosensitizer as a function of the delivery vehicle, Photochem. Photobiol., 60(2),<br />

154–159, 1994.<br />

63. Soncin, M., Polo, L., Reddi, E., Jori, G., Kenney, M. E., Cheng, G., and Rodgers, M. A., Effect of the<br />

delivery system on the biodistribution of Ge(IV) octabutoxy-phthalocyanines in tumour-bearing mice,<br />

Cancer Lett., 89(1), 101–106, 1995.<br />

64. Soncin, M., Polo, L., Reddi, E., Jori, G., Kenney, M. E., Cheng, G., and Rodgers, M. A., Effect of axial<br />

ligation and delivery system on the tumour-localising and -photosensitising properties of Ge(IV)octabutoxy-phthalocyanines,<br />

Br. J. Cancer, 71(4), 727–732, 1995.<br />

65. Wohrle, D., Muller, S., Shopova, M., Mantareva, V., Spassova, G., Vietri, F., Ricchelli, F., and Jori,<br />

G., Effect of delivery system on the pharmacokinetic and phototherapeutic properties of bis(methyloxyethyleneoxy)<br />

silicon-phthalocyanine in tumor-bearing mice, J. Photochem. Photobiol. B,<br />

50(2–3), 124–128, 1999.<br />

66. The basics of boron neutron capture therapy, http://web.mit.edu/nrl/www/bnct/info/description/<br />

description.html, date accessed, Nov. 27, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Nanoemulsion Formulations for Tumor-Targeted Delivery 739<br />

67. Shawer, M., Greenspan, P., OIe, S., and Lu, D. R., VLDL-resembling phospholipid-submicron emulsion<br />

for cholesterol-based drug targeting, J. Pharm. Sci., 91(6), 1405–1413, 2002.<br />

68. Miyamoto, M., Hirano, K., Ichikawa, H., Fukumori, Y., Akine, Y., and Tokuuye, K., Biodistribution<br />

of gadolinium incorporated in lipid emulsions intraperitoneally administered for neutron-capture<br />

therapy with tumor-bearing hamsters, Biol. Pharm. Bull., 22(12), 1331–1340, 1999.<br />

69. Miyamoto, M., Hirano, K., Ichikawa, H., Fukumori, Y., Akine, Y., and Tokuuye, K., Preparation of<br />

Gadolinium-containing emulsions stabilized with phosphatidylcholine surfactant mixtures for neutron<br />

capture therapy, Chem. Pharm. Bull. (Tokyo), 47(2), 203–208, 1999.<br />

70. Watanabe, T., Ichikawa, H., and Fukumori, Y., Tumor accumulation of gadolinium in lipid-nanoparticles<br />

intravenously injected for neutron-capture therapy of cancer, Eur. J. Pharm. Biopharm., 54(2),<br />

119–124, 2002.<br />

71. Teicher, B. A., Use of perfluorochemical emulsions in cancer therapy, Biomater. Artif. Cells Immobil.<br />

Biotechnol., 20(2–4), 875–882, 1992.<br />

72. Teicher, B. A., Herman, T. S., and Frei, E., Perfluorochemical emulsions: Oxygen breathing in<br />

radiation sensitization and chemotherapy modulation, Important Adv. Oncol., 39–59, 1992.<br />

73. Teicher, B. A., Holden, S. A., Ara, G., Ha, C. S., Herman, T. S., and Northey, D., A new concentrated<br />

perfluorochemical emulsion and carbogen breathing as an adjuvant to treatment with antitumor<br />

alkylating agents, J. Cancer Res. Clin. Oncol., 118(7), 509–514, 1992.<br />

74. Rockwell, S., Irvin, C. G., and Nierenburg, M., Effect of a perfluorochemical emulsion on the<br />

development of artificial lung metastases in mice, Clin. Exp. Metastasis, 4(1), 45–50, 1986.<br />

75. Sisk, J., Imaging cancer one nanometer at a time, http://www.radiologytoday.net/imagingcancer.htm,<br />

accessed date, Sept. 06, 2005.<br />

76. Schmieder, A. H., Winter, P. M., Caruthers, S. D., Harris, T. D., Williams, T. A., Allen, J. S., and<br />

Lacy, E. K., Molecular MR imaging of melanoma angiogenesis with alphanubeta-3-targeted paramagnetic<br />

nanoparticles, Magn. Reson. Med., 53(3), 621–627, 2005.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


36<br />

CONTENTS<br />

Solid Lipid Nanoparticles for<br />

Anti-Tumor Drug Delivery<br />

Ho Lun Wong, Yongqiang Li, Reina Bendayan,<br />

Mike Andrew Rauth, and Xiao Yu Wu<br />

36.1 Introduction ....................................................................................................................... 742<br />

36.2 Composition and Structure of SLN .................................................................................. 743<br />

36.2.1 Lipids Used in SLN ............................................................................................ 744<br />

36.2.2 Surfactants Used in SLN..................................................................................... 747<br />

36.2.3 Other Agents Used in SLN ................................................................................. 748<br />

36.2.4 Types, Proposed Structure, and Morphology of SLN ........................................ 748<br />

36.2.4.1 Classical SLN ..................................................................................... 748<br />

36.2.4.2 Nanostructured Lipid Carrier (NLC) ................................................. 749<br />

36.2.4.3 Polymer–Lipid Hybrid Nanoparticles (PLN) .................................... 749<br />

36.3 Manufacturing Methods .................................................................................................... 751<br />

36.3.1 High Pressure Homogenization .......................................................................... 751<br />

36.3.2 Microemulsion..................................................................................................... 751<br />

36.3.3 Solvent Evaporation ............................................................................................ 753<br />

36.3.4 w/o/w Double Emulsion Method ........................................................................ 753<br />

36.4 Characteristic Properties of SLN ...................................................................................... 753<br />

36.4.1 Drug Loading ...................................................................................................... 753<br />

36.4.2 Drug Release ....................................................................................................... 754<br />

36.4.3 Stability of SLN .................................................................................................. 755<br />

36.5 SLN for Anti-Tumor Drug Delivery................................................................................. 756<br />

36.5.1 Cancer Chemotherapy ......................................................................................... 756<br />

36.5.2 Delivery of Cytotoxic Compounds Using SLN.................................................. 757<br />

36.5.3 Poorly Water-Soluble Compounds ..................................................................... 759<br />

36.5.3.1 General Principle................................................................................ 759<br />

36.5.3.2 Topoisomerase I Inhibitors (Camptothecin, Irinothecan) ................. 759<br />

36.5.3.3 Paclitaxel ............................................................................................ 760<br />

36.5.3.4 Other Poorly Water-Soluble Anti-Tumor Compounds<br />

(Etoposide, All-Trans Retinoic Acid, Butyric Acid)......................... 761<br />

36.5.4 Water-Soluble Ionic Salts ................................................................................... 761<br />

36.5.4.1 General Principle................................................................................ 761<br />

36.5.4.2 Doxorubicin and Idarubicin ............................................................... 761<br />

36.5.4.3 Water-Soluble, Non-Ionic Drug Molecule ........................................ 762<br />

36.5.4.3.1 General Principle ........................................................... 762<br />

36.5.4.3.2 Fluorouracil .................................................................... 763<br />

36.5.4.3.3 Overall Remarks on Cytotoxic Drug Delivery ............. 763<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

741


742<br />

36.5.5 Delivery of Chemosensitizing Agents Using SLN............................................. 764<br />

36.5.5.1 Multidrug Resistance and Chemosensitizers ..................................... 764<br />

36.5.5.2 Rationales for Delivery of Chemosensitizers Using SLN................. 764<br />

36.5.5.3 Current Use of SLN for Chemosensitizer Delivery .......................... 765<br />

36.5.5.3.1 Verapamil and Quinidine............................................... 765<br />

36.5.5.3.2 Elacridar (GG918) ......................................................... 765<br />

36.5.5.3.3 Overall Remarks on Chemosensitizer Delivery ............ 766<br />

36.5.6 Effect of SLN Formulation on In Vitro Anti-Tumor Toxicity .......................... 766<br />

36.5.6.1 In Vivo Efficacy of SLN Delivered Antineoplastic Agents .............. 768<br />

36.5.6.2 Improved Pharmacokinetics and Drug Distribution .......................... 768<br />

36.6 Conclusions and Future Perspectives................................................................................ 770<br />

References .................................................................................................................................... 771<br />

36.1 INTRODUCTION<br />

Nanotechnology for Cancer Therapy<br />

Solid lipid nanoparticles (SLN) are colloidal particles of a lipid matrix that is solid at body<br />

temperature. Since their first introduction <strong>by</strong> Müller et al. in 1991, 1 SLN have attracted increasing<br />

interest as a carrier system for therapeutic and cosmetic applications. 2–7 As a drug delivery system,<br />

SLN have been investigated in the last ten years for pharmacological and dermatological formulation<br />

development. They can be administered through a number of routes including parenteral,<br />

peroral, dermal and rectal. 2,4,6,7 Improved bioavailability and targeting capacity have been<br />

observed 8,9 and enhanced cytotoxicity against multidrug resistant cancer cells 10,11 have been<br />

evidenced when SLN are used as the delivery vehicles.<br />

SLN have been proposed as an alternative to other controlled drug delivery systems (CDDS) such<br />

as lipid emulsion, liposome, and polymeric nanoparticles as a result of their several advantages. For<br />

instance, in comparison to lipid emulsion, the solid lipid matrix of SLN makes sustained drug release<br />

possible. The solid lipids also immobilize drug molecules, there<strong>by</strong> protecting the labile and sensitive<br />

drugs from coalescence and degradation, 12 and reduce drug leakage that are commonly seen in many<br />

other CDDS such as liposomes. Compared with some polymeric nanoparticles, SLN are generally<br />

less toxic because physiological and biocompatible lipids are utilized. Meanwhile, all of the less toxic<br />

surfactants that have been applied to other CDDS are equally applicable for SLN preparation. Other<br />

appealing features of SLN include the feasibility for mass production, 13–16 flexibility in sterilization, 4<br />

and avoidance of organic solvents in a typical SLN preparation process. It should be noted that SLN<br />

are also a versatile formulation. Both lipophilic and hydrophilic compounds can be encapsulated and<br />

delivered <strong>by</strong> SLN with modification in the formulation.<br />

The aforementioned useful qualities of SLN make them particularly attractive for the delivery<br />

of cancer chemotherapeutic agents. Anti-tumor drugs, especially the cytotoxic compounds that are<br />

used in conventional chemotherapy, are unique when compared to other classes of drugs in a<br />

number of areas such as the strong toxicity that is typical of cytotoxic drugs often compromises<br />

their therapeutic effects; poor specificity of their drug action in general; and the frequent occurrence<br />

of drug resistance during chemotherapeutic treatment. These issues may all at least be partly<br />

tackled <strong>by</strong> delivering anti-tumor compounds with a suitable drug carrier system. SLN is potentially<br />

a valuable choice for this purpose. The favorable physicochemical characteristics, controlled<br />

release kinetics, and site-specific drug delivery of SLN mean most of the anti-tumor drugs can<br />

be efficiently loaded and delivered to the cancerous tissues at reasonable rates while keeping the<br />

healthy tissues relatively unexposed to these highly toxic agents 17,18 In addition, many compounds<br />

that can be used to reduce drug resistance in cancer cells are also lipophilic (to be discussed later).<br />

SLN may also be used for the delivery of these compounds to further improve the effectiveness of<br />

chemotherapy of cancer normally resistant to cytotoxic drugs.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 743<br />

In order to fully optimize the performance of SLN for anti-tumor drug delivery, it is critical to<br />

gain some understanding of their fundamental aspects. This starts from knowing their main<br />

constituents that define the physicochemical properties of the SLN system they form, including<br />

the particle size, surface charge, and lipid crystallinity. These, in turn, have a strong impact on the<br />

SLN performance. Drug loading, release kinetics, tumor targeting ability, and pharmacokinetics<br />

in vivo may all be affected. For example, drug encapsulation <strong>by</strong> SLN is inversely correlated to the<br />

degree of crystallinity of lipids that is related to the choice of lipids. 19–21 Surfactants are applied to<br />

stabilize the colloids 22 to delay lipid crystallization 23,24 and modulate the degradation rate of<br />

lipid. 25,26 The tumor targeting capacity of SLN derives from their particle size, colloidal<br />

morphology, and surface modification. The leaky vasculature and poor lymphatic drainage of a<br />

tumor generally facilitate preferential penetration <strong>by</strong> and retention of particulate matters of small<br />

size. 27 The submicron size of SLN allows them to passively target tumor tissues. Further manipulation<br />

of the in vivo distribution of SLN can be achieved <strong>by</strong> modifying their surface charges or<br />

hydrophobicity. All in all, <strong>by</strong> understanding the relationship among the SLN constituents, the<br />

physicochemical properties, and particle behaviors, the anti-cancer performance of SLN can<br />

be optimized. 28,29<br />

SLN are not without their limitations. In fact, some drawbacks of SLN such as the low loading<br />

capacity for hydrophilic drugs, biphasic release kinetics, and formulation stability issues are still<br />

barriers against their popular use. 30,31 Fortunately, this class of drug carriers is also rapidly evolving.<br />

A number of strategies have been developed to overcome these limitations. Some of these<br />

strategies lead to new generations of SLN. In this chapter, an overall review of the fundamental<br />

aspects of SLN, the strategies employed for efficient encapsulation and controlled release of various<br />

anti-cancer drugs <strong>by</strong> SLN, and their current use for anti-tumor drug delivery is presented. It is<br />

believed that the information described here will contribute to the design and development of new<br />

SLN formulations for improved delivery of anti-tumor compounds.<br />

36.2 COMPOSITION AND STRUCTURE OF SLN<br />

The typical ingredients of SLN consist of solid lipid(s), surfactant(s) (plus co-surfactants if<br />

required), and incorporated active ingredients. To date, the lipids applied in the preparation of<br />

SLN include the following categories based on their structural diversity: fatty acids with different<br />

lengths of hydrocarbon chain; fatty esters; fatty alcohol; glycerides with different structures; and<br />

mixtures of glyceryl esters. The frequently used anionic and non-ionic lipids and their structures are<br />

listed in Table 36.1a and the cationic lipids in Table 36.1b. Waxes such as Cutina CP (cetylpalmitate),<br />

solid paraffin, and beewax are also sometimes used for SLN preparation. The surfactants<br />

used include amphoteric, ionic, and non-ionic ones (Table 36.2). Although ionic surfactants such as<br />

sodium dodecyl sulphate will allow the SLN to possess narrow size distribution and better stability,<br />

some of them are associated with undesirable toxicity. In addition to the conventional lipid-based<br />

SLN, the synthesized para-acyl-calix[4]arene-based SLN was also recently developed. 32<br />

SLN have the inherent advantage to poorly encapsulate water-soluble agents because of the<br />

lipophilicity of solid lipids. However, to incorporate hydrophilic drugs into SLN, some additional<br />

measures have to be taken. Counterions such as polymers and esters (organic salts) are required to<br />

neutralize the charge of the drugs (both listed in Table 36.2). This may increase the apparent<br />

partition coefficient of the selected drug in the lipid phase and enhance its loading. The encapsulation<br />

of specific hydrophilic anti-tumor compounds is further discussed in Section 36.5.2.2 and<br />

Section 36.5.2.3.<br />

According to the drug incorporation mechanisms, SLN-based systems can be classified into the<br />

following categories, i.e., classical SLN, polymer–lipid hybrid SLN (PLN), nanostructured lipid<br />

carriers (NLC), and lipid-drug conjugate (LDC). Their differences are summarized in Table 36.3.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


744<br />

TABLE 36.1a<br />

Structures of Lipids Used in the Preparation of Solid Lipid Nanoparticles (SLN)<br />

C<br />

H 3<br />

C<br />

H 2<br />

O<br />

HC O<br />

C<br />

H 2<br />

C H2<br />

Chemical Structure Name of Lipid<br />

n<br />

OC<br />

OC C H2<br />

COOH Dodecanoic acid (nZ10)<br />

Myristic acid (nZ12)<br />

Palmitic acid (nZ14)<br />

Stearic acid (nZ16)<br />

C<br />

H 2<br />

C O OC C<br />

H H2<br />

2<br />

O<br />

HC OH<br />

C<br />

H 2<br />

C<br />

H 2<br />

OH<br />

O<br />

HC OH<br />

H 2<br />

C<br />

H 2<br />

C<br />

2<br />

OH<br />

O<br />

HC OH<br />

C<br />

H 2<br />

C<br />

H OH<br />

O<br />

HC O<br />

C<br />

H 2<br />

H 31 C 15<br />

OH<br />

OC C H2<br />

OC<br />

C<br />

H 2<br />

OC C H2<br />

OC<br />

CH 3<br />

CH 3<br />

36.2.1 LIPIDS USED IN SLN<br />

n<br />

n<br />

n<br />

n<br />

CH 3<br />

CH 3<br />

OH<br />

C C<br />

n m<br />

H<br />

C<br />

H 2<br />

OC C H 2<br />

O C 16 H 33<br />

CH 3<br />

H 2<br />

CH 3<br />

Caprylate triglycerides (nZ6)<br />

Caprate triglycerides (nZ8)<br />

Trilaurin (nZ10)<br />

Tripalmitin (nZ14)<br />

Tristearin (nZ16)<br />

Tribehenin (nZ20)<br />

Glyceryl monostearate (nZ16)<br />

Glyceryl hydroxystearate (nZ10, mZ5)<br />

n Glyceryl behenate (mono-, nZ20)<br />

n<br />

CH 3<br />

n Glyceryl behenate (di-, nZ20)<br />

CH 3<br />

O Cetylpalmitate<br />

Nanotechnology for Cancer Therapy<br />

The choice of lipids is critical for SLN to achieve the desired drug loading capacity, stability, and<br />

sustained release behavior. Different lipid phases lead to different apparent partition coefficients<br />

and different loading capacities for the same drug. Polymorphism of lipids also affects the properties<br />

of a SLN system. In addition, hydrophobicity of lipids varies with the balance of the<br />

hydrophobic and hydrophilic functional groups on the lipid molecules. For instance, a fatty acid<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 745<br />

TABLE 36.1b<br />

Cationic Lipids Used for Preparation of Cationic SLN<br />

Stearylamine (SA)<br />

N,N-di-(b-steaorylethyl)-N,N-dimethylammonium chloride (EQ1)<br />

Benzalkonium chloride (BA)<br />

Cetrimide (CTAB)<br />

Dimethyldioctadecylammonium bromide (DDAB)<br />

N-[1-(2,3-dioleoyloxy)propyl-N,N,N-trimethylammonium chloride (DOTAP)<br />

Chloroquine phosphate (CQ)<br />

Cetylpyridinium chloride (CPC)<br />

Dimethyldioctadecylammonium bromide (DDAB)<br />

TABLE 36.2<br />

A List of Surfactants, Co-Surfactants, Counterions, and Surface-Modifying Agents<br />

Commonly Used in the Formulation of SLN-Based Systems<br />

Surfactants and Co-Surfactants<br />

Amphoteric surfactants Non-ionic surfactants<br />

Egg phosphatidylcholine (EggPC) Brij78<br />

Egg lecithin (Lipoid E80) Poloxamer 188<br />

Soy phosphatidylcholine (SP) Poloxamer 407<br />

(Epikuron 200, 95% SP) Poloxamine 908<br />

(Lipoid S100) Polyglycerol methyl glucose distearate<br />

(Lipoid S75, 68% SP) (Tego care 450)<br />

(Lipoid S75, 68% SP) Solutol HS15<br />

(Phospholipon 90G, 90%) Span 85<br />

Trehalose<br />

Ionic surfactants Tween 20<br />

Sodium cholate Tween 80<br />

Sodium cocoamphoacetate Tyloxapol<br />

Sodium dodecyl sulfate (SDS)<br />

Sodium glycocholate Co-surfactants<br />

Sodium oleate Butanol<br />

Sodium taurocholate Butyric acid<br />

Sodium taurodeoxycholate<br />

Counterions<br />

Organic salts Ionic polymers<br />

Monodecylphosphate Dextran sulfate sodium salt<br />

Monohexadecylphosphate Hydrolyzed, polymerized epoxidized soy<br />

bean oil (HPESO)<br />

Mono-octylphosphate<br />

Sodium hexadecylphosphate<br />

Surface-modifying Agents For long-circulating SLN<br />

Dipalmitoylphosphatidyl-ethanolamine conjugated with polyethylene glycol 2000 (DPPE-PEG 2000)<br />

Distearoyl-phosphatidylethanolamine–N-poly(ethylene glycol) 2000 (DSPE–PEG)<br />

stearic acid-PEG 2000<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


746<br />

TABLE 36.3<br />

Categories of Solid Lipid Nanoparticles (SLN)<br />

Nanotechnology for Cancer Therapy<br />

SLN 1. Solid lipid matrix as carrier<br />

2. Lipophilic drug or hydrophilic drug–ester complex is molecularly incorporated into<br />

the solid lipid matrix<br />

NLC 1. Binary mixture of solid lipid and spatially different liquid oil as carrier<br />

2. Lipophilic drug or hydrophilic drug–ester complex is molecularly incorporated into<br />

the solid lipid matrix<br />

PLN 1. Solid lipid as carrier<br />

2. Hydrophilic drug–polymer complex is incorporated into the solid lipid matrix<br />

LDC 1. Hydrophilic drug–lipid conjugate as the homogeneous matrix<br />

becomes more hydrophobic with the increase in the length of its hydrocarbon chain. A fatty ester is<br />

more hydrophobic than the fatty acid of the same chain length because the hydrophilic carboxyl<br />

group of the fatty acid is replaced <strong>by</strong> the more hydrophobic ester group. Triglycerides are more<br />

hydrophobic than mono- and di-glycerides because all three hydrophilic hydroxyl groups are<br />

substituted <strong>by</strong> fatty ester. The lipid composition essentially determines the overall hydrophobicity<br />

of SLN.<br />

Thus far, there has not been a definite rule available for choice of lipids for SLN preparation<br />

though some empirical guidelines are proposed. Measurement of solubility of a given drug was<br />

suggested for oral drug delivery <strong>by</strong> SLN. 33 Empirical equations were used to mathematically<br />

predict the partition of a drug between a lipid or oil phase and an aqueous phase based on physical<br />

interactions of the solute (drug) and solvents (lipid or aqueous phase). For parenteral SLN formulations,<br />

certain requirements need to be met in the selection of lipids such as the suitability of<br />

producing particulates in a submicron range, sterilization <strong>by</strong> autoclaving, long-term storage stability<br />

either in dispersion or after lyophilization, and acceptable toxicity. 34<br />

In general, high drug loading in SLN can only be obtained when the drug has a high solubility in<br />

the lipid melt or a high partition coefficient between the molten lipid and aqueous phase. Based on<br />

the general principle of like-dissolve-like, Li et al. proposed a method to rationally screen the lipid<br />

carrier for SLN to obtain high drug loading capacity <strong>by</strong> a combination of the theoretical calculations<br />

and experimental partition coefficient measurements. 35 The theoretical calculations include<br />

total solubility parameter, polarity, and mixing enthalpy. This method has already been used for the<br />

screening of 16 lipids from different categories, and it has shown good predictive power.<br />

The occurrence of lipid polymorphism is a phenomenon fairly unique to SLN as compared to<br />

other lipid formulations. Lipid polymorphism is referred to the multiple crystalline forms of solid<br />

lipids. One of these polymorphic forms, usually the one that forms perfect crystalline lattice, is<br />

more thermodynamically stable. For example, triglyceride has three forms, i.e., a-form, b-form,<br />

and b 0 -form. Among them, only the b-form is stable. 36 The relatively less stable or meta-stable<br />

forms would eventually transform to the stable polymorphic form, leading to a number of challenges<br />

in the development of SLN formulations. Drug molecules are primarily loaded into SLN in<br />

the defects of the lattices of solid lipids. When the lipid molecules are converted from the metastable<br />

forms into the stable form, they become more orderly packed, and some of the defects in the<br />

lattices disappear. As a result, drug expulsion from the lipid core to the particle surface may occur,<br />

contributing to high initial burst release and drug leakage during storage.<br />

Therefore, one additional criterion for choosing lipids for SLN formulation is the tendency of<br />

these lipids to form perfect crystals, or more precisely, their polymorphic transition rates. Once<br />

again, there is no strict rule to follow. However, in general, lipids with longer fatty acid chains have<br />

a slower transition rate than those with shorter chains. 22 SLN made from waxes as lipid matrix<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 747<br />

exhibited significant drug expulsion phenomenon because of the more crystalline structure<br />

although wax-based SLN are usually more physically stable. 19 In addition to the type of lipids,<br />

some other factors may also affect the lipid crystallinity, including the storage condition and SLN<br />

production methods. 37–39 For example, rapid cooling process will be beneficial to maintain the lipid<br />

matrix in a meta-stable form. 37 The choice of surfactant and co-surfactant plays a crucial role as<br />

well. 24 The crystallization rate of Dynasan 114-based SLN was shown much faster when poloxamer<br />

was used as the surfactant than those stabilized <strong>by</strong> the sodium salt of cholic acid. 40 An increased<br />

amount of the meta-stable form of stearic acid was also found when a high concentration of<br />

co-surfactant such as butanol was included in the SLN preparation. 25<br />

To avoid the aforementioned crystallinity and polymorphism problems, lipid oils, e.g., Miglyol<br />

812 (caprylic/capric triglycerides) or oleic acid, were incorporated into solid lipids or a binary<br />

mixture of physically incompatible solid lipids was used, attempting to disrupt the crystallinity of<br />

the solid lipid matrix. The resulting modified form of SLN is often referred to as nanostructured<br />

lipid carrier (NLC) (see more description in Section 36.2.4.2). 41–43 This method may also provide<br />

additional advantages such as increased drug payload.<br />

Rather than avoiding the complexity caused <strong>by</strong> lipid polymorphism and crystallinity, some<br />

researchers attempted to exploit this phenomenon, and they turned it into a useful feature. For<br />

instance, the polymorphic conversion of lipid matrix from the meta-stable b 0 form to the stable b<br />

form, as confirmed <strong>by</strong> form wide-angle x-ray scattering pattern, altered drug release rate, 44<br />

suggesting that <strong>by</strong> careful design, release rate of SLN can be modulated <strong>by</strong> controlling the<br />

polymorphic interconversion.<br />

Cationic SLN formulations containing positively charged lipids were developed for gene<br />

delivery 45 as the positive surface charge may enhance in vivo transfection efficiency of genes.<br />

Some of the catioic lipids are listed in Table 36.1b. From the point of view of cytotoxicity, twotailed<br />

cationic lipids are usually preferable to the one-tailed cationic lipids (i.e., CPC and CTAB)<br />

and cationic surfactants.<br />

36.2.2 SURFACTANTS USED IN SLN<br />

The main functions of surfactants in SLN are to disperse the melt lipid into aqueous phase in the<br />

SLN preparation process and then to stabilize the nanoparticles after cooling. A surfactant molecule<br />

has a hydrophilic head and lipophilic tail to reduce the surface tension between two phases. The<br />

hydrophile–lipophile balance (HLB) value of surfactants represents the relative proportion of the<br />

hydrophilic and lipophilic parts of the molecule.<br />

As illustrated in Table 36.2, surfactants applied in SLN include four categories: anionic, e.g.,<br />

sodium stearate and sodium dodecyl sulfate; cationic, dodecyltrimethyl ammonium bromide<br />

(DTAB); non-ionic, e.g., members of the Pluronics w or Tweens w family; and amphoteric, e.g.,<br />

egg lecithin. In general, ionic surfactants employ the electrostatic approach, and non-ionic surfactants<br />

rely on steric repulsion to stabilize particles. Among them, most non-ionic surfactants consist<br />

of a hydrophobic moiety (i.e., hydrocarbon chain) and a hydrophilic part (i.e., ethylene oxide<br />

group). Amphoteric surfactants possess both positively and negatively charged groups and<br />

exhibit cationic surfactant’s feature at low pH and anionic at high pH.<br />

Several factors should be considered when choosing surfactants for SLN preparation, namely<br />

HLB value, route of administration, toxicity, and the effect on lipid modification and particle size.<br />

Surfactants with HLB values in the range of 8–18 are suitable for the preparation of oil-in-water<br />

dispersion. Non-ionic surfactants are preferred for oral and parenteral preparations over ionic ones<br />

because of their lower toxicity and irritancy. In general, the order of the toxicity of surfactants is:<br />

cationic O anionic O non-ionic O amphoteric.<br />

One important factor in the choice of a surfactant for SLN should be considered: different<br />

surfactants have variable influences on the in vivo biodegradation of the lipid matrix. For instance,<br />

non-ionic surfactants are more effective for inhibiting the degradation of lipid matrix in vivo.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


748<br />

Poly(ethylene oxide) (PEO) chains of non-ionic surfactants provided steric hindrance of the<br />

anchoring of lipase/co-lipase complex. Therefore, in vivo degradation rate of SLN may be tailored<br />

<strong>by</strong> adjusting the density of PEO chains on the particle surface. For SLN made of the same lipid but<br />

different non-ionic surfactants, the inhibitory activity against lipid degradation is in the order:<br />

cholic acid, sodium salt (NaCh) ! lipoid E80 ! Tween w 80 ! poloxamer 407. 25,40 Another<br />

study shows that Dynasan 114-based SLN have the fastest degradation rate when stabilized <strong>by</strong><br />

lecithin and NaCh; intermediate rate using Tween w 80; and the slowest rate using poloxamer 407. 26<br />

Mixing NaCh and poloxamer does not affect the degradation rate.<br />

Finally, the choice of surfactants and co-surfactants also affects the particle size of SLN. SLN<br />

made of the same lipids may have different sizes because of the use of different surfactants. 25 Zhang et<br />

al. prepared two types of long-circulating SLN using Brij w 78 or Pluronic w F68 as the surfactant, and<br />

they found that different surfactants led to the SLNs with totally different particle size. 46 Bocca et al.<br />

showed that the average diameters of SLN systems produced <strong>by</strong> using glycocholate as the co-surfactant<br />

were larger than those using taurocholate. 47 The authors attributed this phenomenon to the<br />

difference in pKa values of the co-surfactants, i.e., 4.4 for glycocholate and 1.4 for taurocholate.<br />

36.2.3 OTHER AGENTS USED IN SLN<br />

In addition to lipids and surfactants (and sometimes co-surfactants), in some modified forms of<br />

SLN, other agents may also be used. For example, for the encapsulation of cationic, water-soluble<br />

compounds, a source of counterions is needed that includes organic anions and anionic polymers<br />

(see Table 36.2).<br />

SLN can also be surface-modified. When a drug carrier was coated with hydrophilic polymers<br />

such as poloxamers, poloxamines, or poly(ethylene glycol) (PEG), this so-called stealth or longcirculating<br />

carrier may reduce the capture <strong>by</strong> reticuloendothelial system (RES), and it stays longer<br />

in systemic circulation. 48,49 For a detailed description of the stealth technique, refer to other<br />

chapters in the book that deal with long-circulating nanoplaforms. This technique has been<br />

equally valuable to minimize clearance of SLN <strong>by</strong> phagocytosis. 47 Because of the high surface<br />

hydrophobicity of SLN, the hydrophilic coating agents are usually pre-conjugated to lipid moieties<br />

to form amphiphilic molecules (see Table 36.2) so their attachment on to the particle surface can be<br />

more secure. For the effects of SLN surface modification on the pharmacokinetics and biodistribution<br />

of anti-tumor agents, refer to Section 36.5.6.<br />

36.2.4 TYPES, PROPOSED STRUCTURE, AND MORPHOLOGY OF SLN<br />

The structure and morphology of SLN are affected <strong>by</strong> several factors such as the physicochemical<br />

characteristics of the SLN ingredients and the production method adopted. To date, there has been<br />

no concrete data revealing the detailed structure of SLN. Most SLN researchers believe that SLN<br />

are probably spherical because a spherical subject has the smallest possible surface area-to-volume<br />

ratio. This may minimize the surface tension between the lipids and the aqueous phase, there<strong>by</strong><br />

leading to thermodynamic stability. Nevertheless, it should be noted that the platelet shape of SLN<br />

and NLC was also claimed <strong>by</strong> some groups, and they have provided evidence as well. 50,51<br />

36.2.4.1 Classical SLN<br />

Nanotechnology for Cancer Therapy<br />

Based on the difference of melting points (MP) between drug and lipid matrix together with the<br />

consideration of the release kinetics of SLN, Mühlen et al. proposed three hypothetical models for<br />

classical SLN: drug-enriched core, drug-enriched shell, and solid solution 52 as illustrated in<br />

Figure 36.1. Atomic force microscopy (AFM) and DSC studies of prednisolone-loaded SLN<br />

showed that the particles had an inner crystalline core (hard) and an outside amorphous layer<br />

(soft). 53 The core-enriched model is based on the fact that the compound with a higher MP will<br />

precipitate first during cooling. In contrast, if the drug has a lower MP than the lipid matrix,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 749<br />

(a) (b)<br />

(d) (e)<br />

FIGURE 36.1 Proposed structures of solid lipid nanoparticles (SLN): (a) drug-enriched core model; (b) drugenriched<br />

shell model; and (c) drug molecularly dispersed model. (Adopted from zur Mühlen, A., Schwarz, C.,<br />

and Mehnert, W., Eur. J. Pharm. Biopharm., 45, 149–155, 1998.); and nanostructured lipid carrier (NLC): (d)<br />

micro-compartment model; (e) liquid oil molecularly dispersed model.<br />

a drug-enriched shell model will fit the particles. If the drug has a similar MP to the lipid matrix, a<br />

solid solution model will best describe the particle. Although there is no formal study on the<br />

structure of LDC, it should be similar to classical SLN with the loaded drug evenly distributed<br />

within the lipid matrix because they are directly bonded to the lipid molecules.<br />

36.2.4.2 Nanostructured Lipid Carrier (NLC)<br />

To overcome the difficulties of classical SLN, NLC was introduced. In this case, the lipid matrix is<br />

composed of the binary mixture of a solid lipid and the medium chain triglycerides or liquid oil.<br />

BasedonthecompositionofthelipidmatrixofNLC,Jenningetal.proposedtwodrug<br />

incorporation models 54 as demonstrated in Figure 36.1. In the first model, liquid oil is molecularly<br />

dispersed within the solid lipid matrix when the concentration of the liquid oil is below its solubility<br />

in the lipid. In the second model, liquid oil is distributed in the solid lipid in droplet form. Physically<br />

separate phases (solid, liquid) coexist in this model.<br />

36.2.4.3 Polymer–Lipid Hybrid Nanoparticles (PLN)<br />

In order to deliver the hydrophilic drugs with a high drug loading capacity and simultaneously<br />

control the release kinetics of SLN, a new variation of SLN, PLN was recently developed. 55,56 The<br />

schematic illustration of proposed the formation mechanism and structure of PLN is presented in<br />

Figure 36.2. An ionic drug, e.g., doxorubicin HCl, forms complex with a counter ionic polymer,<br />

e.g., dextran sulfate that then partitions into the lipid phase because of higher hydrophobicity,<br />

enabling incorporation of water soluble drugs in the lipid nanoparticles. PLN have a spherical<br />

Water-soluble<br />

counter polymer<br />

+<br />

Water-soluble<br />

ionic drug<br />

Drug-polymer<br />

complex<br />

(c)<br />

+ lipid<br />

Drug-loaded polymer-lipid<br />

Hybrid nanoparticles (PLN)<br />

FIGURE 36.2 Proposed formation mechanism and structure of polymer–lipid hybrid nanoparticles (PLN).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


750<br />

FIGURE 36.3 Transmission electron microscope (TEM) image of polymer–lipid hybrid nanoparticles (PLN)<br />

loaded with doxorubicin. (From Wong et al., Pharm. Res., in press.)<br />

morphology as shown in Figure 36.3. Modest initial burst release and subsequent sustained release<br />

profiles have been observed using this system. However, if the drug–polymer complex is uniformed<br />

distributed in the solid lipid matrix as shown in Figure 36.2 or just attached to the surface of<br />

particles needs further verification.<br />

Li et al. applied differential scanning calorimetry (DSC) and powder x-ray diffraction (PXRD)<br />

crystallography to examine crystallinity and crystal structures of PLN made of verapamil HCl,<br />

dextran sulfate, and dodecanoic acid. 35 The DSC and the PXRD results (Figure 36.4) showed that<br />

Intensity<br />

2 10 20 30<br />

2 theta, degrees<br />

40 50<br />

FIGURE 36.4 Overlaid powder x-ray diffraction crystallographs of (1) verapamil HCl (VRP); (2) dodecanoic<br />

acid; (3) dextran sulfate sodium salt (DS); (4) physical mixture of verapamil HCl and dextran sulfate sodium<br />

salt; (5) VRP–DS complex; and (6) VRP–DS complex incorporated dodecanoic acid. (From Li et al., Pharm.<br />

Res., 2005.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 751<br />

the crystal structure of the lipid remained, but its melting temperature decreased a bit as the drug<br />

became amorphous in the complex form. This finding suggests that the presence of drug–polymer<br />

complex in PLN may be similar to the liquid oils that reduce the lipid crystallinity in NLC.<br />

Nevertheless, the exact significance of the finding will require further investigations.<br />

36.3 MANUFACTURING METHODS<br />

Currently, many techniques are available for SLN preparation. These include high pressure homogenization,<br />

57 microemulsion, 58 and solvent diffusion. 59,60 The more commonly used methods will<br />

be briefly described in this chapter. For detailed descriptions of SLN preparation, please refer to the<br />

up-to-date reviews 30,38,57–60 where the optimization of the processing variables is also discussed.<br />

36.3.1 HIGH PRESSURE HOMOGENIZATION<br />

High pressure homogenization methods include hot and cold approaches. The main advantages of<br />

these methods include narrow size distribution of nanoparticle produced and the possibility of<br />

scale-up production. The cold approach minimizes the exposure of drug to heat (although brief<br />

period of heating is still required) and is more suitable for labile drugs sensitive to heating.<br />

However, this approach requires more tedious steps, and it results in particles of more variable<br />

sizes. The general scheme of high pressure homogenization method is shown in Figure 36.5 (hot<br />

method) and Figure 36.6 (cold method).<br />

36.3.2 MICROEMULSION<br />

A warm microemulsion is prepared under stirring before it is dispersed in a large volume of cold<br />

water at 2–48C. The typical volume ratio of emulsion to cold water is in the range of 1–20. The<br />

procedure of microemulsion method for SLN preparation is illustrated in Figure 36.5.<br />

Hot homogenization<br />

method<br />

High-speed stirring to<br />

form coarse emulsion<br />

High pressure homogenization<br />

at a temperature above the M.P.<br />

of lipid<br />

Cool down<br />

Melt lipid and<br />

disperse drug in lipid<br />

Mix the molten lipid and<br />

drug with hot surfactant<br />

solution<br />

Solid lipid nanoparticles<br />

Microemulsion<br />

method<br />

Gentle stirring, high surfactant/<br />

cosurfactant concentration leads<br />

to microemulsion<br />

Dispersing into excess<br />

cold water under stirring<br />

FIGURE 36.5 Schematic procedures of hot homogenization and microemulsion for SLN production.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


752<br />

Cold homogenization method<br />

Melt lipid, disperse<br />

drug in lipid melt<br />

Cool to solidify lipid/drug mixture<br />

Ground into fine powder<br />

under liquid nitrogen<br />

Dispering the lipid/drug powder<br />

into a cold surfactant solution<br />

High-speed stirring to<br />

form coarse emulsion<br />

High pressure homogenization at<br />

low temperature<br />

Solid lipid nanoparticles<br />

FIGURE 36.6 Schematic procedure of cold homogenization for SLN production.<br />

Using this method, it is noted that the particle size, size distribution, and stability of SLN is<br />

correlated with the properties of microemulsion such as the lipids and surfactants applied, the<br />

concentration of surfactant, and the drug/lipid ratio.<br />

Solvent evaporation method<br />

Lipid and drug dissolved in water-immiscible<br />

organic solvents, i.e., cyclohexane or toluene<br />

Mix with the aqueous surfactant<br />

solution<br />

High-speed stirring to produce<br />

coarse emulsion<br />

High pressure homogenization<br />

Vacuum evaporation of organic solvents<br />

Solid lipid nanoparticles<br />

FIGURE 36.7 Schematic procedure of solvent evaporation method for SLN production.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 753<br />

36.3.3 SOLVENT EVAPORATION<br />

As demonstrated in Figure 36.7, the lipid is dissolved in water-immiscible organic solvent first, and<br />

then the hydrophobic drug is added. Dispersing the drug–lipid–organic solvent mixture into an<br />

aqueous solution of emulsifiers produces an o/w emulsion. SLN are obtained <strong>by</strong> removing the<br />

organic solvent with evaporation under low pressure.<br />

This method totally avoids heating. Its main disadvantage is the inclusion of organic solvent in<br />

preparation. Residues of organic solvent may raise the bio-safety concern.<br />

36.3.4 W/O/W DOUBLE EMULSION METHOD<br />

This method is designated for the delivery of hydrophilic drugs <strong>by</strong> using SLN. 61 Highly concentrated<br />

drug solution is dispersed into a lipid phase under vigorous stirring and stabilized <strong>by</strong> a<br />

surfactant to form the primary w/o emulsion. The primary emulsion is subsequently dispersed<br />

into a large volume of aqueous solution containing another kind of surfactant to form w/o/w<br />

emulsion. The nanoparticles are obtained <strong>by</strong> ultrafiltration or solvent evaporation. The detailed<br />

procedure is illustration in Figure 36.8.<br />

36.4 CHARACTERISTIC PROPERTIES OF SLN<br />

36.4.1 DRUG LOADING<br />

Double emulsion method<br />

Dispersing small amount of<br />

hydrophilic drug solution<br />

in molten lipid<br />

Adding stabilizers<br />

Primary w/o emulsion<br />

Dispersing into a large volume of<br />

aqueous phase containing stabilizers<br />

under higher speed stirring<br />

Cooling down<br />

Solid lipid nanoparticles<br />

FIGURE 36.8 Schematic procedure of w/o/w double emulsion method for SLN production.<br />

For classical SLN, NLC, and PLN, the drug loading process involves partitioning and re-partitioning<br />

of the drug between the lipid and aqueous phase during the heating and subsequent cooling<br />

process. No covalent bonding or ionic bonding between the molecules of the drug and lipid occurs.<br />

On the other hand, drug incorporation <strong>by</strong> LDC, another modified form of SLN, involves direct<br />

covalent or ionic bonding of the drug molecules to the lipids.<br />

Two models are currently proposed to describe how drug molecules are incorporated in SLN,<br />

NLC, and PLN. In the first model, drug molecules are accommodated in the flaws of the lipid crystal<br />

lattice. This model has been widely used to interpret the drug loading behaviors and the biphase<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


754<br />

release kinetics of SLN and NLC. In the second model, the physical and chemical compatibilities<br />

(affinity, miscibility) between the drug–polymer complex and the solid lipid phase are used to<br />

predict the drug loading <strong>by</strong> PLN. 35<br />

For water-soluble, ionic anti-tumor compounds, to allow their efficient encapsulation <strong>by</strong> SLN,<br />

three possible strategies have so far been employed. These strategies include mixing the anti-tumor<br />

compound with an organic salt carrying opposite charges and packaging the resulting ion-pairs into<br />

lipids; 62–65 complexing the ionic anti-tumor compound to lipids carrying opposite charges and<br />

directly using the drug–lipid complexes for SLN preparation (i.e., LDC preparation); and<br />

complexing the ionic anti-tumor compound to polymer molecules carrying opposite charges, i.e.,<br />

polyelectrolytes and dispersing these drug–polymer complexes into lipids to form SLN (i.e., PLN<br />

preparation). Figure 36.9 summarizes these three feasible strategies. These strategies are similar in<br />

terms of their underlying principle—to neutralize the charge on the drug with a counter ion to<br />

facilitate drug partitioning into the lipids.<br />

When an organic salt or ionic polymer interacts with the drug molecules of opposite charge, the<br />

resulting complex loses charges and gains in hydrophobicity that favors their incorporation into the<br />

solid lipid matrix. The ionic molar ratio between the organic salt or polymer and the drug plays a<br />

critical role in this process. By employing isothermal titration calorimetry in conjunction with<br />

partition experiments, Li et al. found that at an equal ionic molar ratio of verapamil HCl (VRP)<br />

to dextran sulfate (DS), the VRP-DS complex reached maximum hydrophobicity and partition in<br />

the lipid phase. 35<br />

36.4.2 DRUG RELEASE<br />

(1) Drug + + Ion –<br />

(2) Drug + + Lipid –<br />

Lipid<br />

(3) nDrug + + Polymer n–<br />

Lipid<br />

Drug + –Lipid –<br />

Drug-Ion-SLN<br />

Nanotechnology for Cancer Therapy<br />

[Drug + –Lipid – ]-SLN<br />

[nDrug + – Polymer n– ]-SLN<br />

FIGURE 36.9 Strategies to load cationic, water-soluble anti-tumor compounds into SLN-based systems. (1)<br />

Ion-pair formation method; (2) lipid drug conjugate (LDC); and (3) polymer–lipid hybrid nanoparticles (PLN).<br />

The release kinetics of SLN usually exhibit a biphasic feature: a strong initial burst release (usually<br />

O 50% of the loaded drug in the first several minutes or hours) followed <strong>by</strong> a subsequent prolonged<br />

release. It is a consensus that the release behavior of SLN is determined <strong>by</strong> the physicochemical<br />

properties of the constituents and the morphology of SLN. By correlating the drug release kinetics<br />

to the physicochemical properties of the drug and lipids, the hypothetical structures of SLN can be<br />

deduced (as mentioned in Section 36.2.4.1). In general, sustained release profiles are correlated to<br />

the drug-enriched core model, and both the drug-enriched shell model and the solid solution model<br />

will have fast release kinetics.<br />

High drug loading capacity can be achieved in NLC. However, the existence of the microcompartments<br />

of oil droplet in the solid lipid matrix may compromise the control over the release<br />

kinetics. The control of release can be improved in the NLC model with liquid oil molecularly<br />

dispersed in the solid lipid matrix.<br />

In comparison to SLN and NLC, PLN offer an additional mechanism to regulate the drug release<br />

kinetics. In addition to the regular diffusion control, the need for ion exchange between the drug–<br />

polymer complex and the surrounding medium for the drug to be released may also be exploited for<br />

better drug release control. The synergistic effect of both approaches will possibly lead to reduced<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 755<br />

% Dox released from Dox-PLN<br />

100<br />

initial burst release and a subsequent sustained release. Figure 36.10 presents representative release<br />

profiles of PLN. Moderate initial releases (50% w/w of drug released in first few hours in both dextran<br />

sulfate-based or HPESO-based PLN loaded with doxorubicin) were achieved when 0.05M CaCl 2<br />

was used as the release medium, and sustained releases were observed in the late phase of release.<br />

36.4.3 STABILITY OF SLN<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0 2 4 6 8 10 12 14 16<br />

Time (hours)<br />

Dox-PLN (HPESO)<br />

Dox-PLN (DS)<br />

FIGURE 36.10 Cumulative percent of doxorubicin released versus time for PLN in 0.05 M CaCl 2 solution at<br />

378C. ( ) Dextran sulfate (DS) sodium salt as the counter polymer; and ( ) hydrolyzed polymer of<br />

expoxidized soybean oil (HPESO). (Adapted from Wong et al., Pharm. Res., in press.)<br />

The practical usage of SLN depends on their long-term physical and chemical stability. Typically,<br />

the shelf life of one viable formulation should be at least one year. The criteria for the assessment of<br />

physical stability of SLN include average particle size, zeta potential, and drug leakage during<br />

storage. For labile compounds, it is critical to evaluate their chemical stability as well.<br />

The physicochemical properties of the materials used to prepare SLN such as the lipids, drugs,<br />

and surfactants play a crucial role in determining the stability of SLN. Waxes usually lead to slower<br />

particle growth and aggregation than glycerides. 19 Glycerides that include a fraction of monoglycerides<br />

are usually more stable than those without (e.g., Dynasan 116 O Compritol 888 ATO O<br />

Imwitor 900). Ionic surfactants often stabilize SLN better than non-ionic surfactants.<br />

In addition to the physicochemical properties of the SLN material, the stability of SLN is also<br />

strongly affected <strong>by</strong> the various processes the SLN subject to after their preparation. For instance,<br />

sterilization is strictly needed for parenteral SLN formulation. When compared to most other drug<br />

delivery systems, SLN are surprisingly stable when challenged <strong>by</strong> the harsh conditions<br />

(compressed, saturated steam at 1218C) during autoclaving. Cavalli et al. found no significant<br />

changes occurred for SLN in terms of the particle size, polydispersity (PI) index, and zeta potential<br />

before and after sterilization. 66 However, there are conflicting reports. Venkateswarlu et al. 39<br />

demonstrated an increase of SLN size almost 2–3-fold, a shift of zeta potential from positive to<br />

negative, and a slight drop in drug loading after autoclaving of SLN. The differences between the<br />

two cases are probably attributed to the different lipids, surfactants, and drugs used in their studies.<br />

Lyophilization is another process that has a strong influence on the formulation stability.<br />

Lyophilization, or freeze drying, is necessary to prevent SLN aggregation or hydrolysis in<br />

aqueous medium. In general, carbohydrate cryoprotectants such as trehalose may stabilize a<br />

SLN formulation during lyophilization. 67 Proper freezing velocity and re-dispersion method<br />

(e.g., sonication) are also required. It has been demonstrated that with the optimized parameters<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


756<br />

for lyophilization, the reconstituted SLN are feasible for i.v. injection; 68 although, there are also<br />

studies showing that some deterioration in the particle quality, particularly the particle size, may<br />

occur. 69 Even in these studies, the reconstituted SLN remained sufficiently small in size and stable<br />

for oral administration.<br />

During storage, the stability of SLN will also be affected <strong>by</strong> the energy introduced from outside<br />

sources such as storage temperature, light, and packing materials. 70 Strong light and high temperature<br />

will accelerate the decrease of the particles’ zeta potential and, in turn, induce particle<br />

aggregation because colloids with zeta potential less negative than K15 mV are prone to aggregation.<br />

Some packing materials such as siliconized vials also exhibit statistically significant effects on<br />

the particle size growth in comparison to that packed in untreated glass vials.<br />

Most aforementioned stability data derive from non-biological studies. There have been more<br />

in vivo SLN studies recently on this issue. This is important because the environment of the<br />

systemic circulation is actually harsh on SLN. For example, the electrolytes in the plasma will<br />

lead to the shrink of double layer, the rapid fall off of zeta potential, and the subsequent coagulation<br />

of particles occurs. 71 Particles may also grow in size and subsequently form semi-solid gels through<br />

adding mono-, di-, and trivalent ions. 72 In addition, the biodegradation of the lipid matrix caused <strong>by</strong><br />

enzymes in vivo will accelerate drug release from SLN and expose the agents to the various<br />

degradative systems before they can reach the disease site. In brief, even when performing SLN<br />

studies in vitro, it will be advisable to at least use media that simulate the composition of body fluids<br />

such as Hanker’s buffer salt solution (HBSS) for the experimental works. This may help the<br />

researchers to gain insight into the behaviors of SLN in vivo.<br />

Finally, it must be re-emphasized that the various properties of SLN are strongly affected <strong>by</strong> the<br />

occurrence of lipid polymorphism, and these include stability as well. During storage of SLN, drug<br />

expulsion and subsequent drug leakage may occur as a result of lipid polymorphic transition,<br />

causing re-distribution of drug molecules within particles. For details of polymorphism, refer to<br />

Section 36.2.1. All in all, the factors that accelerate particle size growth as a result of the loss of<br />

surface charge, including high storage temperature and strong light, will also speed up polymorphic<br />

transition rate.<br />

36.5 SLN FOR ANTI-TUMOR DRUG DELIVERY<br />

36.5.1 CANCER CHEMOTHERAPY<br />

Nanotechnology for Cancer Therapy<br />

Although hormonal therapy and immunotherapy are useful for certain types of cancer, chemotherapy<br />

with cytotoxic drugs remains the most established and commonly used form of drug therapy. Sometimes,<br />

chemotherapy with cytotoxic drugs is the only treatment a patient receives. More often,<br />

cytotoxic drugs are used in addition to other modalities (e.g., surgery, radiotherapy) to improve<br />

their effectiveness and prevent cancer recurrence. 73 Until recently, most of the SLN systems for<br />

cancer treatment are designed for cytotoxic drug delivery. Therefore, unless otherwise specified, in<br />

the following discussions the terms cytotoxic drug, anti-cancer drug and anti-tumor drug will be<br />

interchangeably used, even though the latter two terms actually carry broader meaning.<br />

Cytotoxic drugs treat cancers <strong>by</strong> preferentially causing death or arrest in growth of cancer cells.<br />

However, their cytotoxicity is not highly specific for cancer cells. The healthy cells, especially<br />

those rapidly dividing, almost inevitably suffer from the chemotherapeutic treatment as well. 74 The<br />

major goal in cancer chemotherapy is, therefore, to achieve reasonably high drug concentration for<br />

effective cancer cell kill without causing strong toxicity. In addition, clinical resistance to<br />

chemotherapy drug therapy often occurs in cancer management. 75 One prominent form of drug<br />

resistance is related to the expression of ATP-dependent membrane-associated drug transporters in<br />

cancer cells. Cells with over expression of these drug transporters are resistant to a broad range of<br />

structurally diverse anti-cancer compounds, a phenotype often referred to as multi-drug resistance<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 757<br />

(MDR). 75,76 Cytotoxic drugs can be actively transported out of the cancer cells <strong>by</strong> these transporters,<br />

rendering the chemotherapy ineffective even when an adequate dose of drug is administered.<br />

36.5.2 DELIVERY OF CYTOTOXIC COMPOUNDS USING SLN<br />

SLN have been quickly gaining ground as a carrier of anti-tumor drugs. 10,11,55,62,77–90 Some of<br />

these SLN systems have been evaluated for their anti-tumor activities, stabilities, and/or biodistribution<br />

in biological systems such as cultured cell lines and animal models. These systems and<br />

TABLE 36.4<br />

SLN Formulations Used for Delivery of Drugs with Anti-Cancer Properties and a Brief<br />

Summary of Their Works<br />

Drugs <strong>Group</strong>s a<br />

F/C b<br />

Focus of Studies<br />

In Vitro In Vivo P/B c<br />

Stealth Comments/References<br />

Camptotecan Yang Y<br />

Poorly Water-Soluble<br />

Y(mice) Stability study/ 91,92<br />

Cholesteryl Serpe<br />

butyrate<br />

73<br />

Y In colon cancer cell<br />

line/ 73<br />

Etoposide Murthy Y Y Routes of admin<br />

compared/ 93<br />

Irinotecan Unger Y Y Y SN38, irinotecan prodrug<br />

used, stability<br />

studied/ 94,95<br />

Paclitaxel Gasco Y<br />

Lee Y Y Y Paclitaxel prodrug<br />

used/ 97<br />

Muller Y<br />

98<br />

Serpe Y In HT28 colon cancer<br />

cell line/ 73<br />

Zhang Y Y(mice) 2 formulations using<br />

different surfactants<br />

compared/ 99<br />

Retinoic acid Kim Y In various human cancer<br />

cell lines/ 90<br />

Idarubicin Gasco Y<br />

Ionic, water-soluble<br />

Y (rabbit) Duodenal route/ 101<br />

Doxorubicin Gasco Y Y (rat),<br />

(rabbit)<br />

Y<br />

8,100,102<br />

Serpe Y In colon cancer cell<br />

line/ 73<br />

Wu Y Y In MDR breast cancer<br />

cells, cytotoxicchemosensitizer<br />

combination/ 10,60<br />

5-Flurouracil Wang Y<br />

Non-ionic, water-soluble<br />

Y(mice) FudR, derivative of<br />

5-fluouracil used/ 103<br />

a<br />

According to corresponding author.<br />

b<br />

F/C- formulation and characterization.<br />

c<br />

P/B Pharmacokinetics and/or biodistribution studies presented—pharmacokinetics.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


758<br />

related work are listed in Table 36.4. To this point, studies of SLN systems have demonstrated the<br />

following properties of SLN that are desirable for anti-tumor drug delivery:<br />

1. Versatility and modifiability of the carrier that allow encapsulation of cytotoxic agents of<br />

vastly diverse physicochemical properties;<br />

2. Improved anti-cancer drug stability;<br />

3. Capability to load more than one drug in one carrier system that opens up the possibility<br />

of a drug combination delivery (e.g., a cytotoxic agent and a chemosensitizer);<br />

4. Improved in vitro cytotoxicity against cancer cells, including those normally refractory<br />

to chemotherapy;<br />

5. Enhanced drug efficacy in animal models; and<br />

6. Improved pharmacokinetics and in vivo drug distribution.<br />

In the following sections, the use of properties (1) and (2) <strong>by</strong> SLN for lipophilic anti-tumor drug<br />

delivery will be jointly discussed in Section 36.5.2.1, and Section 36.5.2.2 through 36.5.2.5 will<br />

focus on properties (3) to (6), respectively.<br />

Anti-tumor drugs can be categorized based on their solubility properties. In general, the<br />

majority of anti-tumor agents, particularly the cytotoxic compounds, are fairly lipophilic and<br />

poorly water-soluble. 91 There are several exceptions. Some anti-tumor drugs are water-soluble<br />

because their molecules are small and contain multiple hydrophilic groups. Examples of these<br />

water-soluble drugs include 5-fluorouracil (5-FU) (Figure 36.11a, MW Z130.1) and mitomycin-C<br />

(MW Z334.3). In addition, there are some lipophilic anti-tumor molecules that can be<br />

converted into water-soluble, ionic salts because of their ionizable basic functional groups. These<br />

salts are much more commonly used in clinical settings than their corresponding water-insoluble<br />

free bases because they can be conveniently dissolved and administered using standard aqueous<br />

diluents such as 5% dextrose and 0.9% saline. For example, the hydrochloride salts of anthracyclines,<br />

e.g., doxorubicin and idarubicin (Figure 36.11b), and sulfates of alkaloids, e.g., vincristine<br />

and vinblastine, are all commercially available and frequently used in cancer chemotherapy.<br />

As previously discussed, the efficiency and extent of drug incorporation into SLN are strongly<br />

influenced <strong>by</strong> the partition behavior of the drug in a lipid–water two-phase system. The lipophilic,<br />

uncharged anti-tumor drugs may be encapsulated with relative ease using classical SLN preparation<br />

methods. The other two classes (uncharged, hydrophilic compounds and charged, lipophilic<br />

compounds) of water-soluble anti-tumor agents, on the other hand, require new strategies to<br />

enhance their incorporation into SLN. Means to control the release rates of these compounds<br />

F<br />

O<br />

N<br />

H<br />

N<br />

H<br />

O<br />

(a) 5-fluorouracil<br />

R1<br />

O<br />

O<br />

OH<br />

OH<br />

O<br />

R2<br />

OH<br />

H<br />

O O CH3 *NH2 (b) Doxorubicin /Idarubicin<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 36.11 Examples of the molecular structures of two different types of water-soluble anti-cancer<br />

compounds. (a) 5-fluorouracil (5-FU); and (b) doxorubicin (R 1ZOCH 3,R 2ZCH 2OH) or idarubicin (R 1Z<br />

H, R 2ZCH 3). * Basic amine group that is protonated when forming the water-soluble hydrochloride salt of<br />

doxorubicin or idarubicin.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

OH


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 759<br />

from SLN are also necessary because they may rapidly diffuse from the drug carriers into the<br />

surrounding aqueous medium.<br />

36.5.3 POORLY WATER-SOLUBLE COMPOUNDS<br />

36.5.3.1 General Principle<br />

There have been several SLN systems successfully formulated for encapsulation of poorly watersoluble<br />

anti-tumor compounds. These include camptothecin-based topoisomerase I inhibitors,<br />

including camptothecin and irinotecan and their prodrugs; taxanes, mainly paclitaxel; and other<br />

lipophilic compounds with anti-tumor properties. In general, these poorly water-soluble<br />

compounds partition reasonably well in the lipid phase and can be efficiently encapsulated <strong>by</strong><br />

the conventional SLN without much nanoparticle modifications required. These agents, however,<br />

carry other problems such as poor stability and likelihood of precipitation. In fact, each group of the<br />

compounds has its specific problems to be solved before they can be properly delivered using SLN.<br />

36.5.3.2 Topoisomerase I Inhibitors (Camptothecin, Irinothecan)<br />

Camptothecin is a plant alkaloidal compound extracted from Camptotheca acuminata. 92 It serves as<br />

the prototype compound of a relatively new class of anti-tumor agents known as topoisomerase I<br />

inhibitors (e.g., irinotecan, topotecan). The molecules of camptothecin and its related compounds<br />

all possess extensive aromatic structures and are quite lipophilic (Figure 36.12a). 93 The anti-tumor<br />

activity of camptothecin is strongly correlated to the functionality of the lactone ring of the drug<br />

molecule. However, in an aqueous environment, particularly at basic pH, this lactone ring is<br />

vulnerable to hydrolysis, leading to the formation of therapeutically inactive carboxylate. The<br />

poor water solubility and chemical stability of this drug has seriously diminished its practical<br />

value. These two shortcomings can both be overcome <strong>by</strong> the use of SLN for their capabilities to<br />

efficiently load lipophilic compounds and provide a lipid environment to prevent hydrolytic<br />

degradation. Yang et al. 79 described a SLN system for camptothecin delivery. A very high encapsulation<br />

efficiency (99.6%) of camptothecin was achieved. The same study also showed that the<br />

drug mostly remained in its active lactone form until it was released, confirming the protective<br />

effect of SLN against hydrolytic degradation.<br />

Irinotecan (or Camptosar or CPT-11) is a newer member in the topoisomerase I inhibitor<br />

family. It is relatively hydrophilic and stable compared to camptothecin and has been put to clinical<br />

use for colon cancer chemotherapy. However, compared to its parent drug SN-38 (Figure 36.12b),<br />

the anti-tumor activity of irinotecan is nearly a thousand-fold weaker. 94 SN-38 is prevented from<br />

being put to clinical use mainly because of its extreme hydrophobicity. Moreover, hydrolytic<br />

opening of the labile lactone ring of SN-38 and irinotecan is still possible, though less likely. To<br />

N<br />

N<br />

O<br />

N<br />

N<br />

O<br />

O<br />

H3C OH<br />

O<br />

–<br />

O<br />

OH<br />

HO<br />

OH<br />

H3C OH O<br />

(a) (b)<br />

–<br />

CH 2CH 3<br />

N<br />

H 3 C<br />

N<br />

O<br />

O<br />

O<br />

OH<br />

FIGURE 36.12 Molecular structures of two topoisomerase I inhibitors. (a) The hydrolytic degradation of<br />

camptothecin from its active lactone form (left) to its inactive carboxylate form (right). (b) SN-38, a more<br />

potent lipophilic parent drug of irinotecan.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


760<br />

improve its water solubility and chemical stability, SLN formulation of SN-38 was prepared. 81 The<br />

concentrations of SN-38 in the SLN formulation reached as high as 1.0 mg/ml. The lactone ring<br />

stability of the compound was also improved. After 3 h incubation in human serum albumin, the<br />

lactone concentration in SN-38 loaded-SLN remained 80%. This was superior to the retentions of<br />

only 40% lactone for unformulated SN-38 and irinotecan and even more so when compared to free<br />

camptothecin that lost more than 95% of its active lactone form in 90 min.<br />

In brief, SLN are a useful drug carrier for camptothecin-based anti-tumor compounds. Both of<br />

the studies mentioned above have demonstrated the value of SLN for drug solubilization and<br />

preservation of drug stability.<br />

36.5.3.3 Paclitaxel<br />

Nanotechnology for Cancer Therapy<br />

Paclitaxel (Taxol) is an anti-microtubule agent with broad spectrum anti-tumor activities. It is<br />

gaining popularity among oncologists for its effectiveness against several types of malignancies,<br />

even when used alone. 95 Paclitaxel is a poorly water-soluble drug. At present, it is commercially<br />

available as a non-aqueous micellar solution containing a polyoxyethylated castor oil Cremophor<br />

EL and 49.7% dehydrated ethanol. Cremophor EL is known to cause serious hypersensitivity<br />

reactions and nephrotoxicity in human subjects. 96 In order to solubilize paclitaxel without the<br />

need for Cremophor EL, a few SLN formulations have been developed and studied. 62,82–86 In<br />

general, the drug loading in these SLN formulations ranged from 2 to 5%. In these SLN formulations,<br />

surfactants or stabilizers such as Pluronic w F68 (i.e., poloxamer 188), Brij-78, and<br />

phosphatidylcholine were used. In comparison to Cremophor EL, all of these excipients are<br />

commonly included in parenteral formulations with better track records of safety.<br />

Another major issue associated with the conventional paclitaxel formulation is its poor physical<br />

stability. Paclitaxel is routinely diluted to a concentration of 0.6–1.2 mg/ml with saline or dextrose<br />

for clinical use. The dilution can destabilize the formulation and drug precipitation frequently<br />

occurs, necessitating filtration of the drug solution prior to administration. This extra step adds<br />

to the labor cost and may affect the drug concentration. This problem is apparently not found in<br />

paclitaxel-SLN formulations. In one of these studies, 62 the SLN formulation of paclitaxel was even<br />

diluted to nanomolar range using cell culture medium for in vitro cytotoxicity evaluation. There<br />

was no report of drug precipitation.<br />

Drug precipitation also happens when paclitaxel is too quickly discharged from a carrier system<br />

and the concentration of the released drug exceeds the saturation point. One solution is to limit the<br />

rate of drug release. As demonstrated in the paclitaxel-loaded SLN system <strong>by</strong> Cavelli et al., 82 no<br />

drug precipitation was observed because only 0.1% of paclitaxel was released in pseudo-first order<br />

in 2 h.<br />

Besides SLN, researchers have also successfully encapsulated paclitaxel in other drug delivery<br />

systems, particularly those that are lipid emulsion based. 97–99 However, all of these formulations<br />

are prone to drug precipitation during storage. It is believed that paclitaxel has limited solubility in<br />

lipid as well, and the limited solubilities of paclitaxel in both aqueous and lipid phases leads to the<br />

strong tendency of drug precipitation when the formulations are stored for a long term. Stevens<br />

et al. 84 described the use of a lipophilic paclitaxel prodrug (paclitaxel-2 0 -carbonyl-cholesterol) to<br />

overcome this difficulty. Despite the strong lipophilicity of the prodrug, this group was able to<br />

demonstrate over 10% of drug release into bovine serum albumin in a relatively linear fashion<br />

within 2 h. In addition, even without resorting to the use of prodrug, Cavalli et al. showed that their<br />

formulation containing the unaltered paclitaxel was autoclavable at 1218C without a significant<br />

reduction in the amount of the incorporated drug. 82 The same formulation was also stable for over<br />

18 months after lyophilization, and the lyophilized nanoparticles appeared easily dispersable when<br />

cryoprotectant (2% trehalose) was added.<br />

Therefore, the main shortcomings of paclitaxel, including a high risk of hypersensitivity and<br />

frequent drug precipitation, are correctable using SLN. A number of new, more cancer-type-<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 761<br />

specific paclitaxel analogs have been synthesized in the past decade. 100,101 It will be interesting to<br />

see if SLN are equally useful for these compounds.<br />

36.5.3.4 Other Poorly Water-Soluble Anti-Tumor Compounds (Etoposide,<br />

All-Trans Retinoic Acid, Butyric Acid)<br />

Etoposide, all-trans retinoic acid (ATRA), and butyric acid are all lipophilic molecules that have<br />

anti-tumor properties. Out of these three compounds, etoposide (Vepesid, VP-16) has the longest<br />

history in cancer chemotherapy and is still in clinical use for the treatment of several forms of<br />

cancer. Etoposide is a plant alkaloid. It inhibits topoisomerase II and produces reactive derivatives<br />

that can lead to DNA strand break. 102 This drug was successfully encapsulated in SLN made of<br />

tripalmitin. 80 The authors quoted that the particles were stable after four months of preparation with<br />

excellent redispersibility. However, probably because the focus of that study was on the evaluation<br />

of drug biodistribution and in vivo anti-tumor efficacy, few details in this area was revealed. The<br />

SLN system demonstrated an excellent encapsulation efficiency of 98.96% with 4% payload of<br />

etoposide obtained.<br />

Both ATRA and butyric acid remained at the experimental stage in terms of cancer therapy.<br />

ATRA is well documented for its chemical instability. 103 It is sensitive to heat, light, and oxidation,<br />

and it may isomerize into isotretinoin or be oxidized to form inactive products such as all-trans-4oxo.<br />

This compound should be well incorporated into SLN for its high lipophilicity; although, in the<br />

only SLN system documented, the drug payload attempted was merely 2.5 mg ATRA per g of SLN<br />

powder. 77 The chemical stability of ATRA in SLN during storage was evaluated. More than 90% of<br />

the encapsulated drug remained intact after one month of storage at 48C versus less than 60% when<br />

the drug was stored in the forms of methanol sol ution or Tween 80 solution under the same<br />

conditions. Butyric acid is a short chain fatty acid that shows in vitro inhibitory effect against<br />

colon cancer cell growth. However, its in vivo activity is very low because of its rapid metabolism.<br />

99 It was believed that <strong>by</strong> using its prodrug cholesteryl butyrate and <strong>by</strong> delivering this prodrug<br />

using SLN, the anti-tumor activity could be preserved. The prodrug was successfully loaded in<br />

SLN, and the prodrug demonstrated good in vitro activity. 62,104,105 Although the in vivo activity has<br />

not yet been confirmed, the studies provided another example of the value of SLN to protect a<br />

labile chemical.<br />

36.5.4 WATER-SOLUBLE IONIC SALTS<br />

36.5.4.1 General Principle<br />

The anti-tumor drug salts typically have lipophilic molecular structures. The main obstacle against<br />

these salts from efficient loading into SLN is their ionic charges. For details of the drug loading<br />

mechanism of this class of compounds into SLN-based systems, refer to Section 36.4.1. In essence,<br />

it is about neutralizing the charge on the ionic drug salt with a counter ion. It may be argued that<br />

these extra steps can simply be avoided <strong>by</strong> using the free bases of these agents. However, the use of<br />

free base compounds will essentially lead to the same scenario described previously in Section<br />

36.5.1.1. These poorly water-soluble free base compounds will likely be very slowly released, and<br />

the released drug concentrations will not be as high as compared to when the salt forms are used.<br />

The free base compounds may also require dissolution in organic solvents first during the preparation<br />

process to allow even mixing with the lipids. In general, it is preferable to use the watersoluble<br />

salt of an anti-tumor drug for SLN formulation whenever it is available.<br />

36.5.4.2 Doxorubicin and Idarubicin<br />

Although there are quite a few anti-tumor drugs available in ionic salt forms for improved water<br />

solubility, only two of the anthracyclines have been studied for delivery <strong>by</strong> SLN-based<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


762<br />

formulations. These include doxorubicin and idarubicin. These two compounds are similar in<br />

structures except that doxorubicin is more hydrophilic because of the presence of additional<br />

polar substituent groups (Figure 36.11b). Each of these two drug molecules carry an amine<br />

group and can be converted into hydrochloride salts. To facilitate the loading of these salts into<br />

SLN, they need to be made lipid-soluble based on one of the above-mentioned strategies. Currently,<br />

there is still no LDC system designed for anti-tumor drug delivery. The other two strategies<br />

described in Section 36.4.1 (ion-pair, PLN) have both been used and tested.<br />

Gasco’s group used decyl phosphate or hexadecyl phosphate to form ion pairs with doxorubicin<br />

hydrochloride and idarubicin hydrochloride to enhance their loading of into SLN made of stearic<br />

acid and egg lecithin. 87 The ion pair formation resulted in increases in lipophilicity as defined <strong>by</strong><br />

apparent partition coefficients between water and stearic acid at 708C <strong>by</strong> more than 1000-fold for<br />

doxorubicin and more than 300-fold for idarubicin. SLN prepared using these lipophilic ion pairs<br />

carried payloads of doxorubicin and idarubicin up to 7% and 8.4%, respectively. Probably because<br />

of the high lipophilicity, the drug also released very slowly. For both drugs, less than 0.1% drug was<br />

released after 2 h. With these release rates, there lies a theoretical risk that only the cancer cells in<br />

physical contacts with the SLN may be exposed to sufficiently high drug concentrations and<br />

properly treated. In addition, as demonstrated in several studies, chronic exposure to sub-lethal<br />

concentrations of cytotoxic drugs can induce expression of Pgp. 106 Cancer cells that are farther<br />

away from the SLN could possibly develop resistance to the future chemotherapeutic treatment.<br />

Another feasible approach to encapsulate anthracycline salts is to prepare PLN instead. With<br />

the inclusion of dextran sulfate to provide counterions, PLN of doxorubicin hydrochloride were<br />

successfully prepared. 55 The encapsulation efficiency of doxorubicin in the PLN, depending on the<br />

drug payload, was generally over 70% in the presence of dextran sulfate versus approximately 40%<br />

in its absence. More recently, another doxorubicin-loaded PLN formulation that used a more<br />

lipophilic HPESO polymer (hydrolyzed and polymerized epoxidized soybean oil) was formulated.<br />

10 Payload over 6% was achieved. The drug releases from both PLN formulations were<br />

fast compared to ion pair-based SLN (see Figure 36.10). The lower drug encapsulation efficiency<br />

and faster drug release compared to the ion pair-based SLN indicate that unlike ion pair formation,<br />

the drug–polymer complexation probably does not completely neutralize all of the charges on the<br />

polymer molecules. These residual charges might help draw the water molecules into the lipid<br />

matrix to accelerate its disintegration, and they may lead to faster and more complete drug release<br />

from the nanoparticles. If this will lead to more effective cancer therapy still requires<br />

further investigations.<br />

36.5.4.3 Water-Soluble, Non-Ionic Drug Molecule<br />

36.5.4.3.1 General Principle<br />

Nanotechnology for Cancer Therapy<br />

All of the strategies mentioned in Section 36.4.1 rely on neutralization of the charge on the drug<br />

molecules. For small, non-ionic hydrophilic molecules, these strategies cannot be applied. In fact,<br />

researchers have already encountered similar technical challenges in the formulation of SLN for the<br />

treatment of non-cancer diseases. For instance, 3 0 azido-3 0 deoxythymidine (AZT) is a non-ionic<br />

antiviral drug with low molecular weight (MWZ267.24). This drug is too water soluble to be<br />

formulated into SLN. One way to overcome this difficulty is to prepare a lipophilic drug derivative.<br />

In the case of AZT, researchers conjugated a palmitate group to the AZT molecule, and the drug<br />

loading was shown as significantly improved subsequent to the increased lipophilicity. 107<br />

Hydrophilic cytotoxic compounds are relatively few. Both 5-FU and mitomycin-C fall into this<br />

category. Cisplatin and its analog carboplatin can also be considered as water-soluble small drugs.<br />

Until recently, only 5-FU has been attempted for SLN encapsulation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 763<br />

36.5.4.3.2 Fluorouracil<br />

The drug-derivatization strategy was adopted for SLN encapsulation of 5-FU. The water solubility<br />

of 5-FU is 12.2 mg/ml that makes it very difficult to be loaded in SLN. Wang et al. reduced the<br />

water solubility <strong>by</strong> conjugating two octanoyl groups to the 5-FU molecule, 90 resulting in<br />

3 0 ,5 0 -dioctanoyl-5-fluoro-2 0 -deoxyuridine (FuDR). This lipophilic drug derivative could be<br />

loaded into SLN with an encapsulation efficiency at over 90%. This strategy, however, requires<br />

tedious procedures of chemical synthesis. The chemical purity and identity of the final product have<br />

to be confirmed, and the toxicity and efficacy of the derivative also needs to be evaluated. All these<br />

steps demand substantial amounts of work. Drugs like 5-FU can be easily diluted and administered<br />

without the concern for drug precipitation such as in the cases of the lipophilic compounds, this<br />

probably provides very little motivation for the researchers to go through all these tedious<br />

procedures. However, one must keep in mind that SLN provide more than just drug solubilization<br />

and stabilization. They also allow controlled release, and they possibly improve the anti-tumor<br />

activities, side effect profiles, and biodistribution patterns of the loaded drugs. Successful development<br />

of SLN formulations for this class of anti-tumor drugs can still be rewarding.<br />

36.5.4.3.3 Overall Remarks on Cytotoxic Drug Delivery<br />

Cytotoxic anti-tumor drugs are a class of compounds well known for their heterogeneity in terms of<br />

molecular structures and physicochemical properties. Although a number of strategies have been<br />

devised to allow encapsulation of water-soluble salts of cytotoxic compounds, so far, only two<br />

anthracyclines have been successfully encapsulated. It is obvious that there is more to explore in<br />

this area. For low molecular weight and neutral water-soluble compounds, the major limitation of<br />

SLN—inefficient loading of hydrophilic molecules—is still a concern. Until recently, only limited<br />

efforts using the drug-derivatization strategy have been implemented. The various strategies that<br />

may be suitable for encapsulation of the different classes of cytotoxic compounds <strong>by</strong> SLN are<br />

summarized in Table 36.5. Some of the newer SLN-based systems have not been tested for their use<br />

for anti-tumor drug delivery (e.g., NLC and LDC), but based on their general properties, it is<br />

expected that they should be able to adequately handle the suggested drug types. It is expected<br />

that even more well-designed systems will be developed in the near future to allow more specific<br />

and efficient cytotoxic drug delivery.<br />

TABLE 36.5<br />

A Summary of the Possible SLN-Based Systems Currently Available That Can Be Used for<br />

the Delivery of Cytotoxic Drugs and Chemosensitizers<br />

Drug Physicochemical Properties Possible Systems<br />

Cytotoxic drugs Lipophilic, water-insoluble molecules SLN, NLC<br />

Water-soluble, ionic salts Ion-pair-SLN, LDC, PLN<br />

Water-soluble, hydrophilic small molecules SLN or NLC with lipophilic drug derivative<br />

Chemosensitizers Lipophilic, amphiphilic molecules in salt<br />

forms<br />

SLN, NLC similar to cytotoxic drugs<br />

SLN, classical SLN and microemulsion-based SLN; LD, lipid-drug conjugate lipid nanoparticles; PLN, polymer lipid hybrid<br />

nanoparticles.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


764<br />

36.5.5 DELIVERY OF CHEMOSENSITIZING AGENTS USING SLN<br />

36.5.5.1 Multidrug Resistance and Chemosensitizers<br />

As previously mentioned, MDR phenotype in cancer cells presents a significant obstacle to antitumor<br />

therapy at a cellular level. MDR was demonstrated when tumor cells that have been exposed<br />

to one cytotoxic agent develop cross resistance to a broad range of structurally and functionally<br />

unrelated compounds. 108 The cytotoxic drugs that are most frequently associated with MDR are<br />

typically hydrophobic, amphipathic products, including vinca alkaloids (vincristine, vinblastine),<br />

taxanes (paclitaxel, docetaxel), epipodophyllotoxins (etoposide, teniposide), anthracyclines (doxorubicin,<br />

daunorubicin, epirubicin), topotecan, and mitomycin C. 109,110<br />

MDR mediated <strong>by</strong> membrane-bound drug efflux transporters such as P-glycoprotein (Pgp) is<br />

probably the most studied form of drug resistance. 76,111 If MDR as a result of Pgp expression could<br />

be reduced, it would possibly increase the success rate of cancer chemotherapy. One approach is to<br />

use chemicals with Pgp modulatory activity to restore the drug sensitivity of cancer-demonstrating<br />

MDR. These agents, including verapamil, cyclosporin A, quinidine, tamoxifen, and several calmodulin<br />

antagonists, were identified in the 1980s, 110 and they were often referred as chemosensitizers<br />

or multi-drug resistance reversal agents. Newer and more potent analogs of the older chemosensitizers,<br />

including Valspodar (PSC833), Biricodar, and later, the more Pgp-specific synthetic<br />

compounds such as XR9576, 112 LY335979, 113 and Elacridar (GG918/GF120918) 114 were<br />

successively developed.<br />

36.5.5.2 Rationales for Delivery of Chemosensitizers Using SLN<br />

When tested in clinical settings, the earlier chemosensitizers often produced disappointing results,<br />

primarily because of their low affinities for Pgp. High doses of these agents are required to achieve<br />

the desired chemosensitizing effect, and this results in unacceptable toxicity 115 Many of these<br />

chemosensitizers are also substrates for other enzymes and transporters. This further leads to<br />

unpredictable pharmacokinetic interactions when used with other anti-tumor agents. Although<br />

the recently developed chemosensitizers are more potent, less toxic, and more Pgp-specific, pharmacokinetic<br />

interactions still occur 115–117 even when the newest agents such as GG918 were<br />

chosen. 118<br />

Like most Pgp substrates, chemosensitizers are usually hydrophobic molecules. Many chemosensitizers<br />

also carry basic, ionizable groups like doxorubicin. Figure 36.13a and Figure 36.13b<br />

show the molecular structures of two examples of chemosensitizers. These properties allow them to<br />

be incorporated into SLN at high efficiency. By delivering chemosensitizers using SLN, it is hoped<br />

MeO<br />

MeO<br />

CN CH3 C (CH2) 3 N<br />

CH(CH3 ) *<br />

2<br />

CH2CH2 OMe<br />

OMe<br />

OMe<br />

(a) Verapamil (b) GG-918<br />

FIGURE 36.13 Examples of molecular structures of two chemosensitizers. (a) Verapamil; and (b) GG918.<br />

* Basic amine group that is ionizable <strong>by</strong> protonation.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

O<br />

N<br />

H<br />

Nanotechnology for Cancer Therapy<br />

O<br />

N<br />

H<br />

N *<br />

OMe<br />

OMe


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 765<br />

that the risk of pharmacokinetic drug interactions will be further minimized. Moreover, even<br />

without any surface-modifications, SLN are able to improve tumoral drug concentration (to be<br />

discussed later), so the chemosensitizing effects could potentially be improved, and the non-cellular<br />

drug resistance that is contributed <strong>by</strong> subtherapeutic tumoral drug concentration may also<br />

be alleviated.<br />

36.5.5.3 Current Use of SLN for Chemosensitizer Delivery<br />

So far, only a handful of chemosensitizers have been encapsulated <strong>by</strong> SLN. These include cyclosporin-A,<br />

verapamil, quinidine, and GG918. The cyclosporin-A SLN formulation was actually<br />

intended for immunomodulation, the official indication of cyclosporin-A, rather than for drug<br />

resistance reversal. 119 The discussion will, therefore, be focused on the other three compounds.<br />

36.5.5.3.1 Verapamil and Quinidine<br />

Verapamil and quinidine are both lipophilic molecules commonly available in ionic salt forms<br />

(verapamil hydrochloride and quinidine sulfate) that can both partition into a lipophilic phase when<br />

complexed to an anionic polymer. It was determined that the encapsulation efficiencies of the salts<br />

of verapamil and quinidine in lipid nanoparticles were improved from 19.6 to 69.3% and 29.5 to<br />

58.3%, respectively, when dextran sulfate was co-loaded to form PLN. 55 When verapamil and<br />

quinidine were simultaneously loaded, those encapsulation efficiencies slightly dropped to 50.3%<br />

and 55.8%, still significantly higher than without the addition of an anionic polymer. The drug<br />

release rates of both drugs from the dual-loaded PLN system were similar to when the drugs were<br />

separately loaded. The study shows that it is possible to simultaneously load and deliver two<br />

chemosensitizers <strong>by</strong> PLN with only marginal interferences between each other. This could be<br />

therapeutically valuable as it was reported that combinations of two chemosensitizers, even both<br />

at suboptimal concentrations, were able to achieve synergistic or additive effect to overcome MDR<br />

because of Pgp-over expression. 120<br />

PLN dual-loaded with verapamil and doxorubicin were also evaluated. 55 The drug loading<br />

properties and drug release kinetics of the dual-loaded system was similar to the single-drug<br />

systems. Nevertheless, the systemic cytotoxicity of verapamil is very high (LD50Z163 mg/g in<br />

mice). Considering this drug requires a concentration in the range of mM to mM to be effective for<br />

multidrug resistance (MDR) reversal, 121 there is too much risk to put it in clinical use even it is in<br />

encapsulated form. The study, however, reveals the possibility of using SLN-based system for<br />

multiple drug delivery that is critical in terms of cancer chemotherapy as drug combinations are<br />

typically used.<br />

36.5.5.3.2 Elacridar (GG918)<br />

GG918 has strong inhibitory activity on Pgp and breast cancer resistance protein, 122,123 two transporters<br />

that can contribute to cellular drug resistance. It is more potent and safer than<br />

verapamil. 124,125 The use of SLN for simultaneous delivery of GG918 and doxorubicin is still at<br />

experimental stage. 126 It was reported that the two drugs could be efficiently loaded in the same<br />

PLN formulation (co-loaded with HPESO polymer), and the resulting dual-drug system is effective<br />

against Pgp-over expressing cancer cell lines. It is noteworthy that the dual-drug system was also<br />

significantly more effective than when the two drugs were both in solution form or both in separate<br />

single-drug PLN systems. In initial results, 126 the anti-tumor effects of various formulations and<br />

their combinations were as follows:<br />

doxorubicin/GG918-PLN O free doxorubicin/GG918 in solution O doxorubicin-PLN C GG918<br />

solution z doxorubicin solution C GG918-PLN O doxorubicin-PLN C GG918-PLN (a mixture of<br />

two single-drug PLN)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


766<br />

In other words, synergistic effect in drug-resistant cancers may be achieved using this type of<br />

dual-drug SLN (or PLN) system. The drug combination appears to require proper spatial distributions<br />

to optimally function. When the two drugs are separately encapsulated, the synergistic<br />

effect is lost. These findings have laid the foundation for using SLN-based system for combinational<br />

drug chemotherapy.<br />

36.5.5.3.3 Overall Remarks on Chemosensitizer Delivery<br />

The study of SLN systems for the delivery of chemosensitizers is still in its infancy. Only limited<br />

information in this field has been revealed thus far. Given the good lipophilicities of chemosensitizers<br />

in general, it is not surprising to see that the few chemosensitizers studied until now are well<br />

incorporated in PLN. It is expected that other free-base chemosensitizers can be loaded into even<br />

unmodified, classical SLN (See Table 36.5). Formulation of PLN dual-loaded with two chemosensitizers<br />

or a chemosensitizer/cytotoxic agent combination was also shown to be possible. This,<br />

once again, demonstrates the versatility of SLN-based delivery systems.<br />

36.5.6 EFFECT OF SLN FORMULATION ON IN VITRO ANTI-TUMOR TOXICITY<br />

Nanotechnology for Cancer Therapy<br />

The primary goals of cytotoxic drug therapy are to kill as many cancer cells as possible and to<br />

prevent their proliferation. Therefore, it is important to ensure that the drug released from a SLN<br />

system preserves its cytotoxicity, and the drug delivered <strong>by</strong> SLN is at least as cytotoxic to cancer<br />

cells as the conventional free drug solution. These two issues appear identical but are, in fact, quite<br />

different. Drug releases from SLN are rarely 100% complete. In fact, in many SLN formulations to<br />

be discussed later, only fractions of the loaded anti-tumor drugs are releasable. However, these<br />

formulations all demonstrated cancer cytotoxicity comparable or even superior to the corresponding<br />

free drugs. These findings suggest that, in addition to simply unloading the drug<br />

outside the cancer cells and killing the cells with this fraction of free, released drug, the portion<br />

of drug that remains associated with the SLN is still cytotoxic to various extents. SLN quite<br />

possibly have additional mechanisms to deliver this unreleased drug to achieve cancer cell kill<br />

or growth suppression. It is important to evaluate the cytotoxicity of a drug-loaded SLN formulation<br />

as an integral unit and, whenever possible, to also investigate the underlying cytotoxicity<br />

mechanisms that derive from the use of SLN.<br />

Only until recently, a limited number of in vitro cytotoxicity studies on anti-tumor drug-loaded<br />

SLN have been conducted. Miglietta et al. performed trypan blue exclusion assays on breast cancer<br />

cell line MCF-7 and leukemia cell lines (HL60) treated with SLN incorporating doxorubicin or<br />

paclitaxel for 72 h. 83 The results of the assays demonstrated improved cytotoxicity with both<br />

formulations compared to the corresponding free drugs. A similar study using the same methodology<br />

was later carried out to evaluate the cytotoxicity of SLN formulations carrying doxorubicin,<br />

paclitaxel, or cholesteryl butyrate on a colorectal cancer cell line HT-28. 62 SLN of doxorubicin and<br />

cholesteryl butyrate were both significantly more cytotoxic than the corresponding free drugs as<br />

indicated <strong>by</strong> the lower drug concentrations needed for the SLN to achieve 50% cell non-viability<br />

(IC 50 of SLN versus solution: doxorubicin: 81.87 nM versus 126.57 nM, butyrate: 0.3 mM versus<br />

O 0.6 mM;). Paclitaxel SLN were similarly cytotoxic compared to the drug solution. This was<br />

possibly caused <strong>by</strong> the extremely slow release of paclitaxel from the SLN. 82<br />

One interesting finding in Serpe et al.’s study is about the combination use of anti-tumor<br />

drugs. 62 The authors reported that the combination of low concentrations of cholesteryl butyrateloaded<br />

SLN and doxorubicin or paclitaxel in solution exerted synergistic reduction in cancer cell<br />

viability, whereas no such reduction was demonstrated using the same combinations of drugs both<br />

as solution. Although the cytotoxicities appear low and only trypan blue exclusion assays were<br />

employed, the findings nonetheless support the possibility that combination drug therapy involving<br />

SLN formulations could be more effective than the conventional free drug treatment.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 767<br />

Some nanoparticle systems and liposomal formulations have been shown to be effective against<br />

cancer cells with MDR phenotype. 127–130 These systems were all made of lipids or lipophilic<br />

substances, and many of them were coated with non-ionic surfactants. Also with lipid-based<br />

formulations that often use non-ionic surfactants, it is reasonable to expect MDR reversal activities<br />

in at least some SLN systems. In the aforementioned studies, even though HL-60 and HT-28 cell<br />

lines are not specifically drug-resistant cell lines, they are moderately refractory to doxorubicin, and<br />

the SLN formulations were still able to effectively treat them. In another study, PLN dispersed<br />

using Pluronic F68, a non-ionic block co-polymer, were able to enhance the cytotoxicity of doxorubicin<br />

in Pgp-over expressing breast cancer cell line MDA435/LCC6/MDR1. 10 Doxorubicinloaded<br />

PLN also led to significantly stronger reductions in cell viability in trypan blue exclusion<br />

assays than free doxorubicin. In addition, the use of clonogenic assays, a more sensitive method to<br />

evaluate the effectiveness of cancer proliferation suppression, showed a larger than 8-fold increase<br />

in the anti-tumor activity using PLN formulation compared to free doxorubicin (Figure 36.14). It is<br />

interesting to observe that blank particles did not enhance the anti-tumor activity of doxorubicin<br />

solution, and doxorubicin–polymer (Dox–HPESO) was not more effective than free doxorubicin.<br />

The findings of this study further reveal the potential of SLN for drug-resistant cancer treatment,<br />

and they support the earlier claim that encapsulated cytotoxic compound is still effective. In fact, a<br />

cytotoxic compound may need to be encapsulated to achieve that enhanced activity in drugresistant<br />

cancer.<br />

Initial work has been undertaken to delineate the underlying MDR reversal mechanisms <strong>by</strong><br />

SLN. 11 Increased drug uptake and more noticeably, prolonged drug retention, were observed when<br />

doxorubicin-loaded PLN were used instead of free drug. This was further supported <strong>by</strong> fluorescence<br />

microscope images. 11 In Figure 36.15, cancer cells were treated for 2 h and re-incubated in fresh<br />

medium for an additional 2 h. A high level of doxorubicin that has fluorescent property was<br />

intracellularly detected when the drug was delivered in form of PLN. Another fluorescence<br />

microscopy study also revealed that lipids were associated with the loaded drug molecules in the<br />

cells. All these findings suggest that some of the nanoparticles may carry doxorubicin into the cell,<br />

Normalized PE<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

Dox<br />

Dox-HPESO<br />

Dox-PLN<br />

Dox + Blank PLN<br />

0 2 4 6 8 10<br />

Dox concentration (μg/ml)<br />

FIGURE 36.14 Clonogenic assay experiments for the toxicity of Dox human P-glycoprotein overexpressing<br />

MDA435/LCC6/MDR1 breast cancer cell line. The normalized plating efficiencies (normalized PE) after 4 h<br />

exposure to doxorubicin (Dox) solution, Dox-polymer (Dox-HPESO) aggregates, Dox-loaded lipid nanoparticles<br />

(Dox-PLN), or Dox solution C blank PLN are shown. Results are expressed as mean G SD of the<br />

measurements obtained in three separate experiments (nZ6 in each experiment). *, p ! 0.05, significantly<br />

different from Dox solution group. (Adapted from Wong, H. L. et al., Pharm. Res., in press.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


768<br />

FIGURE 36.15 Fluoroescence microscope images demonstrating the effect of a PLN formulation containing<br />

doxorubicin on drug retention and intracellular drug distribution in Pgp-over expressing breast cancer cells<br />

MDA435/LCC6/MDR1. Cells were incubated with nanoparticles containing doxorubicin for 2 h, washed, and<br />

re-incubated in fresh medium for two additional hours. l ex Z540 nm and l em Z590 nm. Magnification of<br />

objective Z100!. Bar represents 20 mm. (Adapted from Wong, H. L. et al., Pharmacol. Exp. Ther., in press.)<br />

possibly <strong>by</strong> endocytosis, <strong>by</strong>passing the membrane-bound drug efflux mechanisms. These drugloaded<br />

nanoparticles could be staying intracellularly to serve as mini-drug depot systems and<br />

chronically inflicting damages to cancer cells.<br />

36.5.6.1 In Vivo Efficacy of SLN Delivered Antineoplastic Agents<br />

In terms of cancer therapy, particularly with regard to solid tumors, in vitro data frequently may<br />

only serve as a good starting point. It is because there are several drug resistance mechanisms,<br />

sometimes referred as non-cellular mechanisms, that exist only in in vivo situations. A treatment<br />

with strong in vitro anti-tumor activity thus maynotbeworkingintumor-bearinganimals.<br />

However, information on the in vivo efficacy of SLN formulations of anti-tumor drugs remains<br />

limited. Most studies that included animal models focused on drug biodistribution. To our knowledge,<br />

only one study with in vivo efficacy component has been published. In this study, mice model<br />

xenografted with HT-28 tumor was used. 62 Tumors treated with the SLN formulation of SN-38 took<br />

longer or comparable time to reach the cut off tumor weight at lower drug dose. The study provided<br />

the first evidence that the ability of SLN to enhance anti-tumor drug effectiveness is not limited in<br />

cultured cells, but may also be applicable to in vivo systems.<br />

36.5.6.2 Improved Pharmacokinetics and Drug Distribution<br />

Nanotechnology for Cancer Therapy<br />

As discussed in Section 36.2.3, SLN can be coated with stealth agents such as PEG to minimize the<br />

clearance <strong>by</strong> RES to achieve extended systemic circulation time. Using stearic acid PEG2000, longcirculating<br />

SLN formulations of doxorubicin and paclitaxel have been formulated. 9,84 A number of<br />

pharmacokinetic and biodistribution studies were based on these SLN systems.<br />

In principle, the unmodified, non-stealth SLN should be rapidly cleared from the systemic<br />

circulation <strong>by</strong> RES. In reality, unmodified SLN were able to remain in the bloodstream for long<br />

times in the animal studies performed. Table 36.6 summarizes the area-under-the-curve (AUC)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 769<br />

TABLE 36.6<br />

A Comparison of the Area-Under-the-Curve Values of Anti-Cancer Drugs Administered in<br />

Solution to Drugs Encapsulated in Unmodified SLN or Long-Circulating SLN<br />

Study <strong>Group</strong> a<br />

Drug b<br />

AUC c<br />

Funaro Yang Wang Zara 99 Zara 02 Zara 02<br />

DOX CTT FuDR DOX DOX IDA<br />

Drug solution 83.7 66 44.5 98.0 47.6 31.4, 108.5 d<br />

Unmodified SLN 814.5 324 138.4 1713.4 123.1 674.2, 458.9<br />

Long-circulating SLN 1121.1 — — — 259.7 e<br />

318.2<br />

433.3<br />

—<br />

a <strong>Group</strong>s named under the first authors of the studied.<br />

b DOX, doxorubicin; CTT, camptothecin; FuDR, 3 0 ,5 0 -dioctanoyl-5-fluoro-2 0 -deoxyuridine; IDA, idarubicin.<br />

c AUC, area under curve, units in min.(g.ml K1 except the studies <strong>by</strong> Yang (h.ng.g K1 ) and Wang (h.mg.ml K1 ).<br />

d First figure and second figure correspond to AUC of IDA after intravenous and duodenal administration, respectively.<br />

e SLN with three different levels of PEG2000-stearic acid.First, second, and third figures correspond to 0.15%, 0.30%, and<br />

0.45% PEG2000 content in lipid microemulsion for SLN preparation.<br />

values of several anti-tumor compounds obtained in pharmacokinetic studies. Drugs in solution,<br />

unmodified SLN, or long-circulating SLN were administered. In all of the studies listed, unmodified<br />

SLN were able to increase the AUC values <strong>by</strong> 3-fold up to over 20-fold, indicating low RES<br />

clearance. Nevertheless, long-circulating SLN still have an advantage over the unmodified SLN.<br />

In the two studies that compared unmodified and long-circulating formulations, the AUC of drugs<br />

delivered <strong>by</strong> the long-circulating SLN were approximately 25% to over 3-fold higher.<br />

In addition to surface modification, the routes of administration also appear to affect the<br />

pharmacokinetics and drug biodistribution. In the etoposide-SLN study, 80 improvements in<br />

tumoral drug accumulation were observed when the etoposide formulations were intraperitoneally<br />

or subcutaneously injected. Subcutaneous administration even led to multiple-fold increases in the<br />

tumor drug concentrations 24 h post-injection. The authors suggested that the slower and progressive<br />

penetration of the nanoparticles (and the loaded drug) from the subcutaneous injection<br />

site into the tumor may result in more favorable patterns of drug distribution. In the study <strong>by</strong> Zara<br />

et al., 88 duodenal administration of idarubicin-loaded SLN also led to a higher AUC than when the<br />

SLN were intravenously administered. It is evident that the biodistribution of an anti-tumor drug<br />

delivered <strong>by</strong> SLN may be further manipulated to achieve the desired therapeutic goal. The manner<br />

<strong>by</strong> which a SLN drug formulation is administered will be a key aspect to consider when designing<br />

animal or clinical studies of SLN for anti-tumor drug delivery.<br />

Regardless of how long an anti-tumor drug can stay in the systemic circulation, the drug has to<br />

finally accumulate in the tumor and avoid the non-cancerous tissues to exercise its therapeutic<br />

effects without causing toxicity to the patient. To confirm the tumor drug accumulation, an in vivo<br />

tumor model is required. Only one SLN study employing tumor-bearing animals has been<br />

performed. 80 Nearly 67% and 30% increases of tumor drug concentrations were measured 1 h<br />

and 24 h post-injection, respectively. This is promising in terms of cancer chemotherapy. More<br />

SLN biodistribution studies involving tumor models will further strengthen the case.<br />

In many of the biodistribution studies of anti-tumor drug-loaded SLN, even without surface<br />

modifications, the drug accumulations in the liver, spleen, and kidney, three RES organs, were<br />

reduced. 8,9,80,88 The role of RES clearance in the biodistribution of SLN may need to be<br />

re-examined. It is also noteworthy to point out the high brain concentrations of SLN delivered<br />

drugs in many of these studies. Cytotoxic drugs do not usually accumulate in the brain tissue<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


770<br />

because of the presence of the blood–brain barrier. This barrier presents a serious challenge to<br />

chemotherapeutic treatment of central nervous system tumors (refer to Chapter 17). The most<br />

prominent mechanism leading to the blood–brain barrier is the high expression of Pgp.<br />

The ability of SLN to carry drugs across the Pgp-rich blood–brain barrier is consistent with the<br />

previously described findings of SLN in Pgp-over expressing cancer cell lines (Section 36.5.3). This<br />

Pgp-<strong>by</strong>passing feature of SLN may be useful for cancer chemotherapy if properly exploited.<br />

In most of the SLN biodistribution studies, the drug concentrations in the heart were significantly<br />

lower using SLN formulations. 8,9,80,88 This is important for anthracycline delivery as this<br />

class of anti-tumor compounds are notorious in causing irreversible cardiotoxicity. By selectively<br />

reducing the anthracycline concentration in the cardiac tissues, higher dose intensity may be given<br />

to achieve better therapeutic outcomes. 131<br />

Overall, SLN are useful for the improvement of the pharmacokinetics and biodistribution<br />

profiles of anti-tumor drugs. Drugs administered tend to stay longer, penetrate the tumor sites<br />

better, and avoid some vulnerable organs. These may be further improved <strong>by</strong> the use of longcirculating<br />

SLN systems.<br />

36.6 CONCLUSIONS AND FUTURE PERSPECTIVES<br />

Nanotechnology for Cancer Therapy<br />

SLN are a promising drug delivery device for rational chemotherapy. After just slightly longer than<br />

a decade of time, SLN already have gone through several stages of development. Although the<br />

ingredients of SLN are relatively simple—basically only lipids and surfactants—there are a diverse<br />

number of parameters involved in the preparation processes that can be manipulated to alter the<br />

SLN properties. These include the physicochemical characteristics of SLN compositions, production<br />

methods, and storage conditions. Significant progresses in the improvement of drug<br />

loading capacity, mass-production, modulation of release kinetics, stability in storage, bioavailability,<br />

and active targeting have been made. In addition, the early, classical SLN that were made of<br />

highly crystallized lipids have quickly evolved into several variations, including microemulsionbased<br />

SLN, NLC, LDC, PLN, and surface-modified SLN. Each of these variations corrects a<br />

specified set of problems commonly encountered during the early stage of SLN development.<br />

Currently not only limited to lipophilic compounds, hydrophilic and charged agents can also be<br />

efficiently encapsulated into SLN-based systems, stored with good stability, and released under a<br />

certain degree of control. This makes SLN an excellent choice for the delivery of anti-tumor drugs,<br />

a highly heterogeneous class of drugs with diverse molecular structures and physical properties and<br />

are usually chemical reactive and strongly toxic.<br />

This relatively new class of delivery systems still has a lot of untapped potential as well as<br />

limitations to overcome. Because of the complexity of such drug delivery systems and complex<br />

interactions existed among various causal factors, the fundamental problems hampering the practical<br />

application of SLN such as drug loading and release mechanisms, the correlation between the<br />

constituents of SLN and the properties of SLN have not been completely solved yet. The current<br />

development of SLN still mainly depends on the trial-and-error approach that leads to occasional<br />

controversial results. It is expected that with better understanding of the structure of SLN and their<br />

mechanisms of drug loading and release, the overall performance of SLN can be further optimized.<br />

The biological aspects of SLN for cancer treatment, including SLN-cell interactions and in vivo<br />

SLN-tumor interactions are areas largely unexplored. In addition, there are a number of new<br />

applications of SLN that have just started developing, e.g., the use of SLN for delivery of drug<br />

combinations such as cytotoxic drug—chemosensitizer. The potentials of SLN for gene therapy and<br />

molecular targeting are also emerging. 84,132 It will not be surprising to witness significant research<br />

progress in the field of SLN in the near future, providing more efficacious, less toxic, and more<br />

specific chemotherapy treatments for cancer.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 771<br />

REFERENCES<br />

1. Müller, R. H., Colloidal carriers for controlled drug delivery and targeting: Modification, characterization<br />

and in vivo distribution, CRC Press, 45–55, 1991.<br />

2. Wissing, S. A., Kayser, O., and Müller, R. H., Solid lipid nanoparticles for parenteral drug delivery,<br />

Adv. Drug Deliv. Rev., 56, 1257–1272, 2004.<br />

3. Müller, R. H. et al., Solid lipid nanoparticles (SLN)-An alternative colloidal carrier system for<br />

controlled drug delivery, Eur. Pharm. Biopharm., 41, 62–69, 1995.<br />

4. Schwarz, C. et al., Solid lipid nanoparticles (SLN) for controlled drug delivery. I. Production,<br />

characterization and sterilization, J. Control. Release, 30, 83–96, 1994.<br />

5. Almeida, A. J., Runge, S., and Müller, R. H., Peptide-loaded solid lipid nanoparticles (SLN):<br />

Influence of production parameters, Int. J. Pharm., 149, 255–265, 1997.<br />

6. Müller, R. H., Redtke, M., and Wissing, S. A., Solid lipid nanoparticles (SLN) and nanostructured<br />

lipid carriers (NLC) in cosmetic and dermatological preparations, Adv. Drug Deliv. Rev., 54,<br />

S131–S155, 2002.<br />

7. Wissing, S. A. and Müller, R. H., Solid lipid nanoparticles as carrier for sunscreens: In vitro release<br />

and in vivo skin penetration, J. Control. Release, 81, 345–355, 2002.<br />

8. Zara, G. P. et al., Intravenous administration to rabbits of non-stealth and stealth doxorubicin-loaded<br />

solid lipid nanoparticles at increasing concentration of stealth agent: Pharmacokinetics and distribution<br />

in brain and other tissues, J. Drug Target., 10, 327–335, 2002.<br />

9. Fundarò, A. et al., Non-stealth and stealth solid lipid nanoparticles (SLN) carrying doxorubicin:<br />

Pharmacokinetics and tissue distribution after I.V. administration to rat, Pharmacol. Res., 42,<br />

337–343, 2000.<br />

10. Wong, H. L. et al., A new polymer–lipid hydrid nanoparticle system increases cytotoxicity of doxorubicin<br />

against multidrug resistant human breast cancer cells, Pharm. Res., 23, 7, 1574–1585, <strong>2006</strong>.<br />

11. Wong, H. et al., A Mechanistic study of enhanced doxorubicin uptake and retention in multidrug<br />

resistant breast cancer cells using a polymer–lipid hybrid nanoparticle (PLN) system, J. Pharmcol.<br />

Exp. Ther., (in press).<br />

12. Westesen, K., Siekmann, B., and Koch, M. H. J., Investigations on the physical state of lipid<br />

nanoparticles <strong>by</strong> synchrotron radiation X-ray diffraction, Int. J. Pharm., 93, 189–199, 1993.<br />

13. Dingler, A. and Gohla, S., Production of solid lipoid nanoparticles (SLN): Scaling up feasibilities,<br />

J. Microencapsul., 19, 11–16, 2002.<br />

14. Marengo, E. et al., Scale-up of the preparation process of solid lipid nanoparticles. Part I, Int.<br />

J. Pharm., 205, 3–13, 2000.<br />

15. Jenning, V., Lippacher, A., and Gohla, S. H., Medium scale production of solid lipid nanoparticles<br />

(SLN) <strong>by</strong> high pressure homogenization, J. Microencapsul., 19, 1–10, 2002.<br />

16. Gohla, S. H. and Dingler, A., Scaling up feasibility of the production of solid lipid nanoparticles<br />

(SLN), Pharmazie, 56, 61–63, 2001.<br />

17. Allen, T. M. and Cullis, P. R., Drug delivery systems: Entering the mainstream, Science, 303,<br />

1818–1822, 2004.<br />

18. Moses, M. A., Brem, H., and Langer, R., Advancing the field of drug delivery: Taking aim at cancer,<br />

Cancer Cell, 4, 337–341, 2003.<br />

19. Jenning, V. and Gohla, S., Comparison of wax and glyceride solid lipid nanoparticles (SLN), Int.<br />

J. Pharm., 196, 219–222, 2000.<br />

20. Jenning, V. and Gohla, S. H., Encapsulation of retinoids in solid lipid nanoparticles (SLN),<br />

J. Microencapsul., 18, 149–158, 2001.<br />

21. Bunjes, H. et al., Incorporation of the model drug ubidecarenone into solid lipid nanoparticles,<br />

Pharm. Res., 18, 287–293, 2001.<br />

22. Aulton, M. E., Pharamaceutics: The Science of Dosage form Design., Churchill Livingstone, Edinburgh.<br />

81–118, 1988.<br />

23. Bunjes, H., Koch, M. H. J., and Westesen, K., Influence of emulsifiers on the crystallization of solid<br />

lipid nanoparticles, J. Pharm. Sci., 92, 1509–1520, 2002.<br />

24. Aquilano, D., Cavalli, R., and Gasco, M. R., Solid lipospheres obtained from hot microemulsions in<br />

the presence of different concentration of cosurfactant: The crystallinizatiion of stearic acid polymorphs,<br />

Thermochim. Acta., 230, 29–37, 1993.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


772<br />

Nanotechnology for Cancer Therapy<br />

25. Olbrich, C. and Müller, R. H., Enzymatic degradation of SLN-effect of surfactant and surfactant<br />

mixtures, Int. J. Pharm., 180, 31–39, 1999.<br />

26. Olbrich, C., Kayser, O., and Müller, R. H., Enzymatic degradation of Dynasan 114 SLN—effect of<br />

surfactants and particle size, J. Nanoparticle Res., 4, 121–129, 2002.<br />

27. Matsumura, Y. and Maeda, H., A new concept for macromolecular therapeutics in cancer<br />

chemotherapy: Mechanism of tumoritropic accumulation of proteins and the antitumor agent<br />

SMANCS, Cancer Res., 6, 193–210, 1986.<br />

28. Müller, R. H., Mader, K., and Gohla, S., Solid lipid nanoparticles (SLN) for controlled drug<br />

delivery—a review of the state of art, Eur. J. Pharm. Biopharm., 50, 161–177, 2000.<br />

29. Kawashima, Y., Nanoparticulate systems for improved drug delivery, Adv. Drug Deliv. Rev., 47,<br />

1–2, 2001.<br />

30. Mehnert, W. and Mader, K., Solid lipid nanoparticles, production, characterization and applications,<br />

Adv. Drug Deliv. Rev., 47, 165–196, 2001.<br />

31. Westesen, K., Bunjes, H., and Koch, M. H. J., Physicochemical characterization of lipid nanoparticles<br />

and evaluation of their drug loading capacity and sustained release potential, J. Control.<br />

Release, 48, 223–236, 1997.<br />

32. Shahgaldian, P. et al., Para-acyl-calix-arene based solid lipid nanoparticles (SLNs): A detailed study<br />

of preparation and stability parameters, Int. J. Pharm., 253, 23–28, 2003.<br />

33. Bummer, P. M., Physical chemical considerations of lipid-based oral drug delivery-solid lipid<br />

nanoparticles, Crit. Rev. Ther. Drug Carrier Syst., 21, 1–20, 2004.<br />

34. Manjunath, K., Reddy, J. S., and Venkateswarlu, V., Solid lipid nanoparticles as drug delivery<br />

systems, Methods Find. Exp. Clin. Pharmacol., 27, 127–144, 2005.<br />

35. Li, Y., Taulier, N., Rauth, A. M., and Wu, X. Y., American Association of Pharmaceutical Scientists<br />

(AAPS) conference. TN, USA, Nov. 7–11, 2005.<br />

36. Chapman, D., The polymorphism of glycerides, Chem. Rev., 62, 433–456, 1961.<br />

37. Bunjes, H., Koch, M. H. J., and Westesen, K., Influence of emulsifiers on the crystallization of solid<br />

lipid nanoparticles, J. Pharm. Sci., 92, 1509–1520, 2003.<br />

38. Hou, D. et al., The production and characteristics of solid lipid nanoparticles (SLNs), Biomaterials,<br />

24, 1781–1785, 2003.<br />

39. Venkateswarlu, V. and Manjunath, K., Preparation, characterization and in vitro release kinetics of<br />

clozapine solid lipid nanoparticles, J. Control. Release, 95, 627–638, 2004.<br />

40. Olbrich, C., Kayser, O., and Müller, R. H., Lipase degradation of Dynasan 114 and 116 solid lipid<br />

nanoparticles (SLN)-effect of surfactants, storage time and crystallinity, Int. J. Pharm., 237,<br />

119–128, 2002.<br />

41. Souto, E. B. et al., Development of a controlled release formulation based on SLN and NLC for<br />

topical clotrimazole delivery, Int. J. Pharm., 278, 71–77, 2004.<br />

42. Müller, R. H., Radtke, M., and Wissing, S. A., Nanostructured lipid matrices for improved microencapsulation<br />

of drugs, Int. J. Pharm., 242, 121–128, 2002.<br />

43. Jenning, V., Thunemann, A. F., and Gohla, S. H., Characterization of a novel solid lipid nanoparticle<br />

carrier system based on binary mixtures of liquid and solid lipids, Int. J. Pharm., 199, 167–177,<br />

2000.<br />

44. Jenning, V., Korting, M. S., and Gohla, S., Vitamin A-loaded solid lipid nanoparticles for topical use:<br />

Drug release properties, J. Control. Release, 66, 115–126, 2000.<br />

45. Tabatt, K. et al., Effect of cationic lipid and matrix lipid composition on solid lipid nanoparticlemediated<br />

gene transfer, Eur. J. Pharm. Biopharm., 57, 155–162, 2004.<br />

46. Chen, D. N. et al., In vitro and in vivo study of two types of ion-circulating solid lipid nanoparticles<br />

containing paclitaxel, Chem. Pharm. Bull., 49, 1444–1447, 2001.<br />

47. Bocca, C. et al., Phagocytic uptake of fluorescent stealth and non-stealth solid lipid nanoparticles, Int.<br />

J. Pharm., 175, 185–193, 1998.<br />

48. Gref, R. et al., The controlled intravenous delivery of drugs using PEG-coated sterically stabilized<br />

nanoparticles, Adv. Drug Deliv. Rev., 16, 215–233, 1995.<br />

49. Illum, L. et al., The organ distribution and circulation time of intravenous injected colloidal carriers<br />

sterically stabilized with a block copolymer–Poloxamine 908, Life Sci., 40, 367–374, 1987.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 773<br />

50. Jores, K., Mehnert, W., and Mader, K., Physicochemical investigations on solid lipid nanoparticles<br />

and on oil-loaded solid lipid nanoparticles: A nuclear magnetic resonance and electron spin resonance<br />

study, Pharm. Res., 20, 1274–1283, 2003.<br />

51. Jores, K. et al., Investigation on the structure of solid lipid nanoparticles (SLN) and oil-loaded solid<br />

lipid nanoparticles <strong>by</strong> photon correlation spectroscopy, field-flow fractionation and transmission<br />

electron microscopy, J. Control. Release, 95, 217–227, 2004.<br />

52. zur Mühlen, A., Schwarz, C., and Mehnert, W., Solid lipid nanoparticles for controlled drug<br />

delivery—drug release and release mechanism, Eur. J. Pharm. Biopharm., 45, 149–155, 1998.<br />

53. zur Mühlen, A. et al., Atomic force microscopy studies of solid lipid nanoparticles, Pharm. Res., 13,<br />

1411–1416, 1996.<br />

54. Jenning, V., Mader, K., and Gohla, S. H., Solid lipid nanoparticles (SLNTM) based on binary<br />

mixtures of liquid and solid lipids: A 1H-NMR study, Int. J. Pharm., 205, 15–21, 2000.<br />

55. Wong, H. et al., Development of solid lipid nanoparticles containing ionically complexed chemometherapeutic<br />

drugs and chemosensitizers, Int. J. Pharm., 93, 1993–2008, 2004.<br />

56. Li, Y., Taulier, N., Rauth, A. M., and Wu, X. Y., Screening of lipid carriers and characterization of<br />

drug–polymer–lipid interactions for the rational design of polymer–lipid hybrid nanoparticles<br />

(PLN), Pharm. Res., (Submitted).<br />

57. Müller, R. H. et al., Solid lipid nanoparticles (SLN)-an alternative colloidal carrier system for<br />

controlled drug delivery, Eur. J. Pharm. Biopharm., 41, 62–69, 1995.<br />

58. Gasco, M. R., Method for producing solid lipid microspheres having a narrow size distribution, US<br />

Patent No. 5250236, 1993.<br />

59. Trotta, M., Debernardi, F., and Caputo, O., Preparation of solid lipid nanoparticles <strong>by</strong> a solvent<br />

emulsification-diffusion technique, Int. J. Pharm., 257, 153–160, 2003.<br />

60. Hu, F. Q. et al., Preparation of solid lipid nanoparticles with clobetasol propionate <strong>by</strong> a novel solvent<br />

diffusion method in aqueous systems and physicochemical characterization, Int. J. Pharm., 239,<br />

121–128, 2002.<br />

61. Cortesi, R. et al., Production of lipospheres as carriers for bioactive compounds, Biomaterials, 23,<br />

2283–2294, 2002.<br />

62. Serpe, L. et al., Cytotoxicity of anticancer drugs incorporated in solid lipid nanoparticles on HT-29<br />

colorectal cancer cell line, Eur. J. Pharm. Biopharm., 58, 673–680, 2004.<br />

63. Cavalli, R. et al., Solid lipid nanoparticles (SLN) as ocular delivery system for tobramycin, Int.<br />

J. Pharm., 238, 241–245, 2002.<br />

64. Cavalli, R. et al., Preparation and evaluation in vitro of colloidal lipospheres containing pilocarpine<br />

as ion pair, Int. J. Pharm., 117, 243–246, 1995.<br />

65. Gasco, M. R., Cavalli, R., and Carlotti, M. E., Timolol in lipospheres, Pharmazie, 47, 119–121,<br />

1992.<br />

66. Cavalli, R. et al., Sterilization and freeze-drying of drug-free and drug loaded solid lipid nanoparticles,<br />

Int. J. Pharm., 148, 47–54, 1997.<br />

67. Shahgaldian, P. et al., A study of the freeze-drying conditions of calixarene based solid lipid<br />

nanoparticles, Eur. J. Pharm. Biopharm., 55, 181–184, 2003.<br />

68. Zimmermann, E., Müller, R. H., and Mäder, K., Influence of different parameters on reconstitution of<br />

lyophilized SLN, Int. J. Pharm., 196, 211–213, 2000.<br />

69. Schwarz, C. and Mehnert, W., Freeze-drying of drug-free and drug-loaded solid lipid nanoparticles<br />

(SLN), Int. J. Pharm., 157, 171–179, 1997.<br />

70. Freitas, C. and Müller, R. H., Effect of light and temperature on zeta potential and physical stability<br />

in solid lipid nanoparticle (SLNTM) dispersions, Int. J. Pharm., 168, 221–229, 1998.<br />

71. Martin, A., Physical Pharmacy: Physical Chemical Principles in the Pharmaceutical Sciences, 4th<br />

ed., Lea & Febiger, Philadelphia, PA pp. 362–388, 1993.<br />

72. Freitas, C. and Müller, R. H., Stability determination of solid lipid nanoparticles (SLN) in aqueous<br />

dispersion after addition of electrolyte, J. Microencapsul., 16, 59–71, 1999.<br />

73. Skeel, R. T., Selection of treatment for the patient with cancer, In Handbook of Cancer<br />

Chemotherapy, Skeel, R. T., Ed., Lippincott, Williams, and Wilkins, Philadelphia, PA, pp. 46–52,<br />

2003.<br />

74. Tipton, J. M., Side effects of cancer chemotherapy, In Handbook of Cancer Chemotherapy, Skeel,<br />

R. T., Ed., Lippincott, Williams, and Wilkins, Philadelphia, PA, pp. 561–580, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


774<br />

Nanotechnology for Cancer Therapy<br />

75. Gottesman, M. M., Mechanisms of cancer drug resistance, Ann. Rev. Med., 53, 615–627, 2002.<br />

76. Endicott, J. A. and Ling, V., The biochemistry of P-glycoprotein-mediated multidrug resistance,<br />

Ann. Rev. Biochem., 58, 137–171, 1989.<br />

77. Lim, S. J. and Kim, C. K., Formulation parameters determining the physicochemical characteristics<br />

of solid lipid nanoparticles loaded with all-trans retinoic acid, Int. J. Pharm., 243, 135–146, 2002.<br />

78. Yang, S. C. et al., Body distribution in mice of intravenously injected camptothecin solid lipid<br />

nanoparticles and targeting effect on brain, J. Control. Release, 59, 299–307, 1999.<br />

79. Yang, S. C. and Zhu, J. B., Preparation and characterization of camptothecan solid lipid nanoparticles,<br />

Drug. Dev. Ind. Pharm., 28, 265–274, 2002.<br />

80. Reddy, L. H. et al., Influence of administration route on tumor uptake and biodistribution of etoposide<br />

loaded solid nanoparticles in Dalton’s lymphoma tumor bearing mice, J. Control. Release, 105,<br />

185–198, 2005.<br />

81. Williams, J. et al., Nanoparticle drug delivery system for intravenous delivery of topoisomerase<br />

inhibitors, J. Control. Release, 91, 167–172, 2003.<br />

82. Cavalli, R., Caputo, O., and Gasco, M. R., Preparation and characterization of solid lipid nanospheres<br />

containing paclitaxel, Eur. J. Pharm. Sci., 10, 305–309, 2000.<br />

83. Miglietta, A. et al., Cellular uptake and cytotoxicity of solid lipid nanospheres (SLN) incorporating<br />

doxorubicin or paclitaxel, Int. J. Pharmaceut., 210, 61–67, 2000.<br />

84. Stevens, P. J., Sekido, M., and Lee, R. J., A folate-receptor-targeted lipid nanoparticle formulation<br />

for a lipophilic paclitaxel prodrug, Pharm. Res., 21, 2153–2157, 2004.<br />

85. Videira, M. A., Almeida, A. J., and Muller, R. H., Formulation and physical stability assessment of<br />

SLN containing paclitaxel. Incorporation of paclitaxel in SLN: Assessment of drug–lipid interaction,<br />

Proc. 3rd World Meeting Pharm. Technol., 453–454, 2000.<br />

86. Chen, D. B. et al., In vitro and in vivo study of two types of long-circulating solid lipid nanoparticles<br />

containing paclitaxel, Chem. Pharm. Bull., 49, 1444–1447, 2001.<br />

87. Cavalli, R., Caputo, O., and Gasco, M. R., Solid lipospheres of doxorubicin and idarubicin, Int.<br />

J. Pharm., 89, R9–R12, 1993.<br />

88. Zara, G. P. et al., Pharmacokinetics and tissue distribution of idarubicin-loaded solid lipid nanoparticles<br />

after duodenal administration to rats, J. Pharm. Sci., 91, 1324–1333, 2002.<br />

89. Zara, G. P. et al., Pharmacokinetics of doxorubicin incorporated in solid lipid nanospheres (SLN),<br />

Pharm. Res., 40, 281–286, 1989.<br />

90. Wang, J. X., Sun, X., and Zhang, Z. R., Enhanced brain targeting <strong>by</strong> synthesis of 3-,5-dioctanoyl-5fluoro-2’-deoxyuridine<br />

and incorporation into solid lipid nanoparticles, Eur. J. Pharm. Biopharm.,<br />

54, 285–290, 2002.<br />

91. Pratt, W. B. et al., The Anticancer Drugs, Oxford University Press, New York.<br />

92. Wall, M. E. et al., Plant antitumor agents. I. The isolation and structure of camptothecin, a novel<br />

alkaloidal leukemia and tumour inhibitor from camptotheca acuminate, J. Am. Chem. Soc., 88,<br />

3888–3890, 1966.<br />

93. Gottlieb, J. A. et al., Preliminary pharmacologic and clinical evaluatin of camptothecin sodium<br />

(NSC-100880), Cancer Chemother. Rep., 54, 461–470, 1970.<br />

94. Takimoto, C. H. and Arbuck, S. G., Topoisomerase I targeting agents: The camptothecins, In Cancer<br />

Chemotherapy and Biotherapy: Principles and Practice, Chabner, B. A. and Longo, D. L., Eds.,<br />

Lippincott, Williams, and Wilkins, Philadelphia, PA, pp. 579–646, 2001.<br />

95. Rowinsky, E. K. and Donehower, R. C., Paclitaxel (taxol), N. Engl. J. Med., 332, 1004–1014, 1995.<br />

96. Singla, A. K., Garg, A., and Aggarwal, D., Paclitaxel and its formulations, Int. J. Pharm., 235,<br />

179–192, 2002.<br />

97. Constantinides, P. P. et al., Formulation development and antitumor activity of a filter-sterilizable<br />

emulsion of paclitaxel, Pharm. Res., 17, 175–182, 2000.<br />

98. Kan, P. et al., Development of nonionic surfactant/phospholipid o/w emulsion as a paclitaxel<br />

delivery system, J. Controll. Release, 58, 271–278, 1999.<br />

99. Bernsdorff, C., Reszka, R., and Winter, R., Interaction of the anticancer agent taxol (paclitaxel) with<br />

phopholipid layers, J. Biomed. Mater. Res., 46, 141–149, 1999.<br />

100. Clarke, S. J. and Rivory, L. P., Clinical pharmacokinetics of docetaxel, Clin. Pharmacokinet., 36,<br />

99–114, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Solid Lipid Nanoparticles for Anti-Tumor Drug Delivery 775<br />

101. Ojima, I. and Geney, R., BMS-184476 Bristol-Myers Squibb, Curr. Opin. Investig. Drugs, 4,<br />

732–736, 2003.<br />

102. Baldwin, E. L. and Osheroff, N., Etoposide, topoisomerase II and cancer, Curr, Med. Chem. Anti-<br />

Cancer Agents, 5, 363–372, 2005.<br />

103. Brisaert, M., Gabriels, M., and Plaizier-Vercammen, J., Investigation of the chemical stability of an<br />

erythromycin-tretionin lotion <strong>by</strong> the use of an optimization system, Int. J. Pharm., 243, 135–146, 2000.<br />

104. Pouillart, P. R., Role of butyric acid and its derivatives in the treatment of colorectal cancer and<br />

hemoglobinopathies, Life Sci., 63, 1739–1760, 1998.<br />

105. Pellizzaro, C. et al., Cholesteryl butyrate in solid lipid nanoparticles as an alternative approach for<br />

butyric acid delivery, Anticancer Res., 19 (5B), 3921–3926, 1999.<br />

106. Hahn, S. M. et al., A multidrug-resistant breast cancer line induced <strong>by</strong> weekly exposure to doxorubicin,<br />

Int. J. Oncol., 14, 273–279, 1999.<br />

107. Heiati, H. et al., Solid lipid nanoparticles as drug carriers. I. Incorporation and retention of the<br />

lipphilic prodrug 3’-azido-3’-deoxythymidine palmitate, Int. J. Pharm., 146, 123–131, 1997.<br />

108. Biedler, J. L. and Riehm, H., Cellular resistance to actinomycin D in Chinese hamster cells in vitro:<br />

Cross-resistance, radioautographic, and cytogenic studies, Cancer Res., 30, 1174–1184, 1970.<br />

109. Ambudkar, S. V., Dey, S., and Hrycyna, C. A., Biochemical, cellular and pharmacological aspects of<br />

the multidrug transporter, Ann. Rev. Pharmacol. Toxicol., 39, 361–398, 1999.<br />

110. Krishna, R. and Mayer, L. D., Multidrug resistance (MDR) in cancer. Mechanisms, reversal using<br />

modulators of MDR and the role of MDR modulators in influencing the pharmacokinetics of anticancer<br />

drugs, Eur. J. Pharm. Sci., 11, 265–283, 2000.<br />

111. Juliano, R. L. and Ling, V., A surface glycoprotein modulating drug permeability in Chinese hamster<br />

ovary cell mutants, Biochim. Acta., 455, 152–162, 1976.<br />

112. Roe, M. et al., Reversal of P-glycoprotein mediated multidrug resistance <strong>by</strong> novel anthranilamide<br />

derivatives, Bioorg. Med. Chem. Lett., 9, 595–600, 1999.<br />

113. Dantzig, A. H. et al., Reversal of P-glycoprotein-mediated multidrug resistance <strong>by</strong> a potent cyclopropyldibenzosuberane<br />

modulator, LY335979, Cancer Res., 56, 4171–4179, 1996.<br />

114. Wallstab, A. et al., Selective inhibition of MDR1 P-glycoprotein-mediated transport <strong>by</strong> the acridone<br />

carboxamide derivative GG918, Br. J. Cancer, 79, 1053–1060, 1999.<br />

115. Robert, J. and Jarry, C., Multidrug resistance reversal agents, J. Med. Chem., 46, 4805–4817,<br />

2003.<br />

116. Thomas, H. and Coley, H. M., Overcoming multidrug resistance in cancer: An update on the clinical<br />

strategy of inhibiting P-glycoprotein, Cancer Control, 10, 159–165, 2003.<br />

117. Wandel, C. et al., P-glycoprotein and cytochrome P450 3A inhibition: dissociation of inhibitory<br />

potencies, Cancer Res., 59, 3944–3948, 1999.<br />

118. Planting, A. S. T. et al., A phase I and pharmacologic study of the MDR converter GF120918 in<br />

combination with doxorubicin in patients with advanced solid tumors, Cancer Chemother. Pharmacol.,<br />

55, 91–999, 2005.<br />

119. Ugazio, E., Cavalli, R., and Gasco, M. R., Incorporation of cyclosporine A in solid lipid nanoparticles<br />

(SLN), Int. J. Pharm., 241, 341–344, 2002.<br />

120. Huang, M. et al., Effect of combinations of suboptimal concentrations of P-glycoprotein blockers on<br />

the proliferation of MDR gene expressing cells, Int. J. Cancer, 65, 389–397, 1996.<br />

121. Belpomme, D., Gauthier, S., and Pujade-Lauraine, E., Verapamil increases the survival of patients<br />

with anthracycline-resistant metastatic breast carcinoma, Ann. Oncol., 11, 1471–1476, 2000.<br />

122. Hyafil, F. et al., In vitro and in vivo reversal of multidrug resistance <strong>by</strong> GF120918, an acridonecarboxamide<br />

derivative, Cancer Res., 53, 4595–4602, 1993.<br />

123. Maliepaard, M. et al., Circumvention of breast cancer resistance protein (BCRP)-mediated resistance<br />

to camptothecins in vitro using non-substrate drugs or the BCRP inhibitor GF120918, Clin. Cancer<br />

Res., 7, 935–941, 2001.<br />

124. den Ouden, D. et al., In vitro effect of GF120918, a novel reversal agent of multidrug resistance, on<br />

acute leukemia and multiple myeloma cells, Leukemia, 10, 1930–1936, 1996.<br />

125. Sparreboom, A. et al., Clinical pharmacokinetics of doxorubicin in combination with GF120918, a<br />

potent inhibitor of MDR1 P-glycoprotein, Anticancer Drugs, 10, 719–728, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


776<br />

Nanotechnology for Cancer Therapy<br />

126. Wong, H. L. et al., Formulation, characterization and in vitro cytotoxicity of solid lipid nanoparticles<br />

carrying polymer-linked doxorubicin and GG918, a novel P-glycoprotein inhibitor, AAPS J., 6,<br />

Abstract T3208, 2004.<br />

127. de Verdiere, A. C. et al., Reversion of multidrug resistance with polyalkylcyanoacrylate nanoparticles:<br />

Towards a mechanism of action, Br. J. Cancer, 76, 198–205, 1997.<br />

128. Soma, C. E. et al., Ability of doxorubicin-loaded nanoparticles to overcome multidrug resistance of<br />

tumor cells after their capture <strong>by</strong> macrophages, Pharm. Res., 16, 1710–1716, 1999.<br />

129. Thierry, A. R. et al., Modulation of doxorubicin resistance in multidrug-resistant cells <strong>by</strong> liposomes,<br />

FASEB J., 7, 572–579, 1993.<br />

130. Romsicki, Y. and Sharom, F. J., The membrane lipid environment modulates drug interactions with<br />

the P-glycoprotein multidrug transporter, Biochemistry, 38, 6887–6896, 1999.<br />

131. Moore, M. J. and Erlichman, C., Pharmacology of anticancer drugs, In The Basic Science of<br />

Oncology, Tannock, I. F. and Hill, R. F., Eds., McGraw-Hill, Toronto, pp. 370–391, 1998.<br />

132. Tabatt, K. et al., Transfection with different colloidal systems: Comparison of solid lipid nanoparticles<br />

and liposomes, J. Control. Release, 97, 321–332, 2004.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


37<br />

CONTENTS<br />

Lipoprotein Nanoparticles<br />

as Delivery Vehicles for<br />

Anti-Cancer Agents<br />

Andras G. Lacko, Maya Nair, and Walter J. McConathy<br />

37.1 Introduction ....................................................................................................................... 777<br />

37.2 Lipoproteins as Vehicles for the Targeted Delivery of Anti-Cancer Agents .................. 777<br />

37.3 LDL as a Delivery Vehicle for Anti-Cancer Agents........................................................ 779<br />

37.4 HDL as a Delivery Vehicle for Anti-Cancer Agents ....................................................... 779<br />

37.5 Conclusions ....................................................................................................................... 781<br />

References.................................................................................................................................... 782<br />

37.1 INTRODUCTION<br />

Nanotechnology and its biomedical applications have exploded in the last several years. Because of<br />

initiatives in Europe 1 and the United States, 2,3 increased research funding and venture capital 4 have<br />

been available to facilitate these developments. Nanotechnology, nanoscience, or nanoengineering<br />

encompass pharmaceutical formulations, devices, chemical processes, or that may be encapsulated<br />

into a compartment of less than 1,000 nm in diameter. This volume is dedicated to the description of<br />

a wide variety of drug delivery devices and their application to cancer therapy.<br />

Although lipoproteins have long been considered as potential vehicles to deliver the “magic<br />

bullet” for targeted cancer chemotherapy, 5,6 they have yet to be considered as prime candidates for<br />

clinical applications as recent reviews on targeted drug delivery 7–9 contain only a limited amount of<br />

discussion on lipoproteins 7 or none at all. 8,9 The purpose of this chapter is to evaluate lipoproteins<br />

as delivery agents for anti-cancer drugs and to discuss the challenges that have so far prevented<br />

clinical applications of lipoprotein-based formulations.<br />

37.2 LIPOPROTEINS AS VEHICLES FOR THE TARGETED DELIVERY<br />

OF ANTI-CANCER AGENTS<br />

Plasma lipoproteins are macromolecular complexes composed of specific protein and lipid components.<br />

The major physiological role of lipoproteins is to transport water-insoluble lipids from<br />

their point of origin to their respective destinations. Lipoproteins contain an outer shell that is made<br />

up of phospholipids, apolipoproteins, and unesterified cholesterol (Figure 37.1a) and an interior<br />

core compartment, accommodating water-insoluble lipids (triacylglycerols and cholesteryl esters).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

777


778<br />

Phospholipid<br />

Unesterified cholesterol<br />

CE CE<br />

CE TG<br />

CE<br />

Apolipoprotein<br />

Cholesteryl<br />

Ester<br />

Triglyceride<br />

(a) (b)<br />

Phospholipid<br />

Unesterified cholesterol<br />

Apolipoprotein<br />

Hydrophobic drug<br />

(paclitaxel)<br />

FIGURE 37.1 Loading of a hydrophobic drug (paclitaxel) into the core region of a generic lipoprotein<br />

(Adapted from Wainberg, P. B., Hosp. Pract., 22, 223–227, 1987.)<br />

This arrangement provides considerable stability to the overall structure of lipoproteins and makes<br />

them particularly suitable for the transport of hydrophobic drugs (Figure 37.1b).<br />

Because of the opportunity to convert hydrophilic anti-cancer agents to highly effective hydrophobic<br />

pro-drugs via chemical modifications, 10–12 there is essentially no structural limit to the type<br />

and kind of pharmaceutical agents that may be transported <strong>by</strong> lipoproteins. Plasma lipoproteins are<br />

spherical particles that range from 5 to over 1,000 nm in diameter (Figure 37.2). Because of their<br />

superior stability and smaller size, the classes of low density lipoproteins (LDL) and high density<br />

lipoproteins (HDL) qualify as drug carrying nanoparticles and will be the subject of this review.<br />

The outstanding features that render lipoproteins suitable carriers of anti-cancer agents may be<br />

summarized as follows:<br />

Density (g/ml)<br />

0.95<br />

1.006<br />

1.02<br />

1.06<br />

1.10<br />

1.20<br />

HDL<br />

5<br />

LDL<br />

IDL<br />

10 20 40 60<br />

Diameter (nm)<br />

FIGURE 37.2 Buoyant density and size distribution of plasma lipoproteins.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

Nanotechnology for Cancer Therapy<br />

VLDL Chylomicrons<br />

Chylomicron<br />

remnants<br />

80 1,000


Lipoprotein Nanoparticles as Delivery Vehicles for Anti-Cancer Agents 779<br />

† Lipoproteins represent natural, biologically compatible components providing enhanced<br />

safety and efficacy.<br />

† Apolipoproteins provide a signal for the receptor mediated uptake of the whole particle<br />

<strong>by</strong> endocytosis 13 or the selective uptake of the drug/core component. 14<br />

† Lipoproteins nearly perfectly exemplify the “magic bullet” concept 6 via the receptormediated<br />

uptake of the encapsulated drug. Because of the over expression of lipoprotein<br />

receptors <strong>by</strong> malignant cells, 15–17 there is an opportunity to attack malignant cells selectively<br />

without substantially impacting normal cells. The targeting potential of<br />

lipoproteins may be further enhanced <strong>by</strong> attaching ligands to the respective lipid or<br />

protein components.<br />

37.3 LDL AS A DELIVERY VEHICLE FOR ANTI-CANCER AGENTS<br />

The application of LDL as a targeted drug delivery vehicle was initiated <strong>by</strong> Krieger et al. 18,19 who<br />

replaced the core of native LDL with cholesteryl linoleate and suggested that hydrophobic<br />

compounds, including drugs, may be incorporated into LDL for diagnostic or therapeutic purposes<br />

in a similar manner. 18 Subsequently, Gal et al. 20 proposed an LDL-based drug delivery system for<br />

treating gynecological neoplasms. To date, numerous studies have been performed with LDL<br />

particles as targeted drug carriers; nevertheless, the enthusiasm for LDL-based drug delivery<br />

systems has been modest as indicated <strong>by</strong> the paucity of approved therapeutic formulations or<br />

ongoing clinical trials at present.<br />

The major challenge in the preparation of reconstituted LDL (rLDL) nanoparticles is the<br />

availability of apolipoprotein B-100 (apoB-100), the major protein component of LDL. 13 In<br />

addition, the high molecular weight of apoB-100 and its tendency to aggregate upon delipidation<br />

present additional difficulties in the preparation of rLDL-based drug formulations. Early attempts to<br />

prepare rLDL/drug complexes were undertaken <strong>by</strong> Lundberg, 21–23 combining egg yolk phosphatidylcholine<br />

(EYPC), a hydrophobic drug (cytotoxic mustard carbamate), and apoB-100. 20 The<br />

resultant nanoparticle had a diameter of about 23 nm and was metabolized <strong>by</strong> fibroblasts in a<br />

manner similar to native LDL. 21 A series of studies <strong>by</strong> Van Berkel et al. developed unilamellar<br />

liposome-like particles, containing apolipoprotein E that were taken up <strong>by</strong> tissues via the LDL<br />

receptor upon injection into rats. 24 Van Berkel et al. also encapsulated a lipophilc derivative of<br />

daunorubicin into these nanoparticles that were subsequently taken up <strong>by</strong> B16 tumors 10 via the<br />

LDL receptor. 25<br />

Maranhao et al. developed an alternate approach to prepare a drug carrying, protein-free<br />

microemulsion (LDE) <strong>by</strong> combining egg yolk phosphatidyl choline, triolein, cholesterol, and<br />

cholesteryl oleate. 26 The LDE particles apparently associated with apoE upon injection into rats<br />

and had clearance rates similar to LDL. These LDE particles were subsequently loaded with either<br />

oleoyl-etoposide 12 or oleoyl-paclitaxel 27 to explore their respective therapeutic potentials. Studies<br />

with oleoyl-paclitaxel in mice showed a nine-fold increase in LD 50 of oleoyl-paclitaxel encapsulated<br />

into LDE vs. the free drug. 27 A novel approach for preparing rLDL/drug complexes was<br />

reported <strong>by</strong> Owens et al. who utilized peptides representing the receptor binding region of apoB<br />

instead of full length apo B. 28 This microemulsion resembling LDL was able to support the growth<br />

of U937 cells similar to native LDL. 29<br />

37.4 HDL AS A DELIVERY VEHICLE FOR ANTI-CANCER AGENTS<br />

The utilization of lipoprotein nanoparticles for the delivery of anti-cancer agents is based on the<br />

hypothesis that rapidly proliferating cells (including cancer cells) have a higher expression of<br />

lipoprotein receptors 15–17 to meet their increased need for cholesterol (new cell membrane<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


780<br />

Nanotechnology for Cancer Therapy<br />

synthesis). This view is consistent with the lower plasma cholesterol levels found in cancer<br />

patients. 30 Clinical studies showed that HDL cholesterol levels were lower in cancer patients<br />

versus normal controls than their LDL-cholesterol. 31 Similar findings were reported for patients<br />

with hematologic malignancies, 32–34 small cell lung cancer, 35 and colorectal adenoma. 36 Recent<br />

findings with the HDL receptors (scavenger receptor B1 [SR-B1], CD36, and LIMPII analogous-1<br />

[CLA-1]) show that breast cancer cells have substantially higher expression of this HDL receptor<br />

than the surrounding normal cells. 37 These findings suggest that targeted delivery of anti-cancer<br />

drugs using an HDL delivery system is likely to be highly effective. 38<br />

The proliferation of adenocarcinoma 39 and other cancer cells 40,41 were shown to be enhanced<br />

<strong>by</strong> HDL or HDL components. In addition, Pussinen et al. have shown that breast cancer cells take<br />

up cholesterol from HDL via the CLA-1 receptor pathway. 42 These findings show that HDL<br />

receptors may play a key role in controlling the proliferative capacity of cancer cells, and they<br />

establish the concept of selective targeting of anti-cancer agents delivered via reconstituted HDL<br />

(rHDL). These findings are consistent with findings from this laboratory, showing that cancer cells<br />

produced substantially stronger immunoblots with an SR-BI antibody than normal fibroblasts. 43 It<br />

has also been shown that the respective uptakes of paclitaxel and cholesteryl esters were highly<br />

correlated (r 2 Z0.88; pZ0.04) when delivered as core components of rHDL nanoparticles, indicating<br />

that the paclitaxel was taken up <strong>by</strong> a receptor-mediated pathway. 43<br />

The delivery of anti-cancer agents via native HDL is exceedingly challenging because of the<br />

labor involved in the large scale isolation of human plasma HDL. Concerns about bio-safety would<br />

likely render the development of the HDL/drug formulations perhaps even more costly than those<br />

employing LDL. 44 So far, there have been only limited attempts to incorporate drugs into native<br />

HDL 45 and none with the ultimate purpose of therapeutic applications. Here, rHDL has proven to be<br />

a much more effective drug delivery vehicle than native HDL.<br />

The nomenclature currently used in the literature is confusing as to what rHDL actually<br />

represents. The terms reconstituted and recombinant are both being used to describe artificially<br />

generated HDL-type nanoparticles although the former designation is becoming more predominant.<br />

The nomenclature is further confounded <strong>by</strong> the naming of rHDL nanoparticles based on the<br />

FIGURE 37.3 Transformation of the discoidal precursor (nascent) HDL to its stable, spherical configuration<br />

via the accumulation of cholesteryl esters, catalyzed <strong>by</strong> lecithin: cholesterol acyltransferase (LCAT). (From<br />

Alexander, E. T., et al., Biochemistry, 44(14), 5409–5419, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Lipoprotein Nanoparticles as Delivery Vehicles for Anti-Cancer Agents 781<br />

different stages in the maturation of the circulating HDL particle (Figure 37.3). 46,47 HDL metabolism<br />

commences <strong>by</strong> the secretion of discoidal, primarily phosphatidyl choline (PC) and<br />

apolipoprotein A-I (apoA-I) containing particles <strong>by</strong> the liver 48 and the intestine. These HDL<br />

precursors are subsequently converted to spherical (mature) HDL 49 <strong>by</strong> the action of lecithin:cholesterol<br />

acyltransferase (LCAT) <strong>by</strong> expanding the particle’s cholesteryl ester core (Figure 37.3).<br />

Scientific reports and patent disclosures have referred to both discoidal and spherical nanoparticles<br />

as “rHDL.” 50–54<br />

For the purpose of formulating drug delivery complexes, spherical nanoparticles have an<br />

advantage over other configurations as their core region can accommodate the drug to be transported<br />

(Figure 37.1). Van Berkel et al. prepared spherical neo-high density lipoproteins as drug<br />

delivery vehicles 54 <strong>by</strong> incorporating the lipophilic derivative of iododeoxyuridine into neo-HDL,<br />

resulting in a nanoparticle resembling circulating apoE-free HDL. 11 Earlier work in this laboratory<br />

yielded stable rHDL/drug nanoparticles composed of Taxol w and the components of circulating<br />

HDL. 43 The paclitaxel from the rHDL/Taxol w nanoparticles was apparently taken up <strong>by</strong> cancer<br />

cells via a receptor mediated mechanism. 43 An enhanced rHDL formulation containing paclitaxel<br />

has been recently developed with increased drug carrying capacity and cytotoxicity toward several<br />

cancer cell lines. 55<br />

37.5 CONCLUSIONS<br />

Lipoprotein-based nanoparticles have the potential to provide a superior performance compared to<br />

most alternative formulations for drug delivery agents because of their natural, biodegradable<br />

ingredients that also help to protect against rapid removal from the circulation <strong>by</strong> the reticuloendothelial<br />

system. 44,56 Additional important issues in favor of lipoproteins nanoparticles are the<br />

biological compatibility of lipoprotein based formulations, 44,56,57 many that have already been<br />

safely injected into human subjects. 58–62 Lipoprotein nanoparticles have substantially smaller<br />

molecular diameter than most liposomal drug preparations, a considerable advantage, as even<br />

smaller liposomes have been considered to have superior pharmacokinetic properties. 63 Another<br />

important issue favoring lipoprotein nanoparticles over alternate drug delivery systems is their<br />

targeting potential via receptor-mediated mechanisms that are over expressed in cancer cells vs.<br />

normal cells. 15,17,37 This type of lipoprotein-based (LDL) drug delivery mechanism has been shown<br />

to be superior in the delivery of omega-3 fatty acids to cancer cells over serum albumin. 64<br />

One of the major advantages of lipoprotein nanoparticles in cancer chemotherapy is that their<br />

targeting potential may be substantially enhanced <strong>by</strong> attaching ligands to their surface lipid or<br />

protein components. Targeted drug delivery is of major interest in cancer chemotherapy where<br />

several monoclonal antibodies have already been employed as targeting agents. 65,66 However, there<br />

have been only limited attempts to explore this drug delivery model; it has essentially unlimited<br />

potential in cancer chemotherapy. Earlier attempts to enhance the targeting of the lipoprotein<br />

nanoparticles was reported <strong>by</strong> Van Berkel et al. <strong>by</strong> using lactosylated derivatives of LDL 67,68 and<br />

neo-HDL. 69 These modified lipoprotein nanoparticles were shown to be preferentially targeted to<br />

parenchymal liver cells 67,68 that provided a model for converting water soluble drugs to pro-drugs<br />

for encapsulaton into lipoproteins 69 and a potential drug delivery vehicle for treating hepatitis B. 70<br />

Additional attempts for enhancing the targeting of lipoprotein nanoparticles involved the attachment<br />

of monoclonal antibodies (MABs) to LDL to target histocompatibility antigens. 45 Recent<br />

studies in this laboratory led to the development of an rHDL particle with folate residues attached to<br />

the apo A-I component. These folate-bearing nanoparticles promoted a substantially enhanced<br />

uptake of paclitaxel <strong>by</strong> OVCAR-3 cells from the modified rHDL/paclitaxel complex.<br />

Despite many potential advantages over other drug delivery models, the application of lipoprotein<br />

nanoparticles for cancer chemotherapy currently lags behind other drug delivery systems.<br />

Although lipoprotein nanoparticles are biologically compatible and offer superb targeting potential<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


782<br />

via receptor-mediated mechanisms, the challenge of obtaining the apolipoprotein starting<br />

material has so far been a major obstacle in the manufacturing of lipoprotein-based<br />

pharmaceutical formulations.<br />

In summary, lipoprotein-based nanoparticles have the potential to perform as superior drug<br />

delivery agents for anti-cancer drugs because of their natural biodegradable components that<br />

protect against their rapid removal from the circulation and their targeting potential via receptors<br />

that are over expressed in cancer cells vs. normal cells.<br />

REFERENCES<br />

Nanotechnology for Cancer Therapy<br />

1. European Science Foundation Policy Briefing, ESF scientific forward look on nanomedicine. www.<br />

esf.org, 2005.<br />

2. U.S. Department of Health and Human Services, National Institutes of Health, and National Cancer<br />

Institute, Cancer NANOTECHNOLOGY plan, A strategic initiative to transform clinical oncology<br />

and basic research through the directed application of nanotechnology, 2004.<br />

3. NCI Alliance for Nanotechnology in Cancer Awards. http://nano.cancer.gov/alliance_awards/<br />

4. Paull, R., Wolfe, J., Hebert, P., and Sinkula, M., Investing in nanotechnology, Nat. Biotechnol., 21,<br />

1144–1147, 2003.<br />

5. Counsell, R. E. and Pohland, R. C., Lipoproteins as potential site-specific delivery systems for<br />

diagnostic and therapeutic agents, J. Med. Chem., 25, 1115–1120, 1982.<br />

6. Sörgel, F., The return of Ehrlich’s ‘Therapia magna sterilisans’ and other Ehrlich concepts. Series of<br />

papers honoring Paul Ehrlich on the occasion of his 150th birthday, Chemotherapy, 50, 6–10, 2004.<br />

7. Dubowchik, G. M. and Walker, M. A., Receptor-mediated and enzyme-dependent targeting of cytotoxic<br />

anticancer drugs, Pharmacol. Ther., 83, 67–123, 1999.<br />

8. Allen, T. M. and Cullis, P. R., Drug delivery systems: Entering the mainstream, Science, 303,<br />

1818–1822, 2004.<br />

9. Guillemard, V. and Saragovi, H. U., Novel approaches for targeted cancer therapy, Curr. Cancer Drug<br />

Targets, 4, 313–326, 2004.<br />

10. Versluis, A. J., Rump, E. T., Rensen, P. C., van Berkel, T. J., and Bijsterbosch, M. K., Stable<br />

incorporation of a lipophilic daunorubicin prodrug into apolipoprotein E-exposing liposomes<br />

induces uptake of prodrug via low-density lipoprotein receptor in vivo, J. Pharmacol. Exp. Ther.,<br />

289, 1–7, 1999.<br />

11. de Vrueh, R. L., Rump, E. T., Sliedregt, L. A., Biessen, E. A., van Berkel, T. J., and Bijsterbosch,<br />

M. K., Synthesis of a lipophilic prodrug of 9-(2-phosphonyl methoxyethyl)adenine (PMEA) and its<br />

incorporation into a hepatocyte-specific lipidic carrier, Pharm. Res., 16, 1179–1185, 1999.<br />

12. Azevedo, C. H., Carvalho, J. P., Valduga, C. J., and Maranhao, R. C., Plasma kinetics and uptake <strong>by</strong><br />

the tumor of a cholesterol-rich microemulsion (LDE) associated to etoposide oleate in patients with<br />

ovarian carcinoma, Gynecol. Oncol., 97, 178–182, 2005.<br />

13. Segrest, J. P., Jones, M. K., De Loof, H., and Dashti, N., Structure of apolipoprotein B-100 in low<br />

density lipoproteins, J. Lipid Res., 42, 1346–1367, 2001.<br />

14. Williams, D. L., Temel, R. E., and Connelly, M. A., Roles of scavenger receptor BI and APO A-I in<br />

selective uptake of HDL cholesterol <strong>by</strong> adrenal cells, Endocr. Res., 26, 639–651, 2000.<br />

15. Tosi, M. R. and Tugnoli, V., Cholesteryl esters in malignancy, Clin. Chim. Acta, 359, 27–45, 2005.<br />

16. Ho, Y. K., Smith, R. G., Brown, M. S., and Goldstein, J. L., Low-density lipoprotein (LDL) receptor<br />

activity in human acute myelogenous leukemia cells, Blood, 52, 1099–1114, 1978.<br />

17. Imachi, H., Murao, K., Sayo, Y., Hosokawa, H., Sato, M., Niimi, M., Kobayashi, S., Miyauchi, A.,<br />

Ishida, T., and Takahara, J., Evidence for a potential role for HDL as an important source of cholesterol<br />

in human adrenocortical tumors via the CLA-1 pathway, Endocr. J., 46, 27–34, 1999.<br />

18. Krieger, M., Brown, M. S., Faust, J. R., and Goldstein, J. L., Replacement of endogenous cholesteryl<br />

esters of low density lipoprotein with exogenous cholesteryl linoleate. Reconstitution of a biologically<br />

active lipoprotein particle, J. Biol. Chem., 253, 4093–4101, 1978.<br />

19. Krieger, M., Goldstein, J. L., and Brown, M. S., Receptor-mediated uptake of low density lipoprotein<br />

reconstituted with 25-hydroxycholesteryl oleate suppresses 3-hydroxy-3-methylglutaryl-coenzyme. A<br />

reductase and inhibits growth of human fibroblasts, Proc. Natl Acad Sci. U.S.A. 75, 5052–5056, 1978.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Lipoprotein Nanoparticles as Delivery Vehicles for Anti-Cancer Agents 783<br />

20. Gal, D., Ohashi, M., MacDonald, P. C., Buchsbaum, H. J., and Simpson, E. R., Low-density lipoprotein<br />

as a potential vehicle for chemotherapeutic agents and radionucleotides in the management of<br />

gynecologic neoplasms, Am. J. Obstet. Gynecol., 139, 877–885, 1981.<br />

21. Lundberg, B., Preparation of drug-low density lipoprotein complexes for delivery of antitumoral<br />

drugs via the low density lipoprotein pathway, Cancer Res., 47, 4105–4108, 1987.<br />

22. Lundberg, B. and Suominen, L., Preparation of biologically active analogs of serum low density<br />

lipoprotein, J. Lipid Res., 25, 550–558, 1984.<br />

23. Lundberg, B., Cytotoxic activity of two new lipophilic steroid nitrogen carbamates incorporated into<br />

low-density lipoprotein, Anticancer Drug Des., 9, 471–476, 1994.<br />

24. Rensen, P. C., Schiffelers, R. M., Versluis, A. J., Bijsterbosch, M. K., Van Kuijk-Meuwissen, M. E.,<br />

and Van Berkel, T. J., Human recombinant apolipoprotein E-enriched liposomes can mimic lowdensity<br />

lipoproteins as carriers for the site-specific delivery of antitumor agents, Mol. Pharmacol., 52,<br />

445–455, 1997.<br />

25. Versluis, A. J., Rensen, P. C., Rump, E. T., Van Berkel, T. J., and Bijsterbosch, M. K., Low-density<br />

lipoprotein receptor-mediated delivery of a lipophilic daunorubicin derivative to B16 tumours in mice<br />

using apolipoprotein E-enriched liposomes, Br. J. Cancer, 78, 1607–1614, 1998.<br />

26. Maranhão, R. C., Cesar, T. B., Pedroso-Mariani, S. R., Hirata, M. H., and Mesquita, C. H., Metabolic<br />

behavior in rats of a nonprotein microemulsion resembling low-density lipoprotein, Lipids, 28,<br />

691–696, 1993.<br />

27. Rodrigues, D. G., Maria, D. A., Fernandes, D. C., Valduga, C. J., Couto, R. D., Ibañez, O. C., and<br />

Maranhão, R. C., Improvement of paclitaxel therapeutic index <strong>by</strong> derivatization and association to a<br />

cholesterol-rich microemulsion: In vitro and in vivo studies, Cancer Chemother. Pharmacol., 55,<br />

565–576, 2005.<br />

28. Owens, M. D., Baillie, G., and Halbert, G. W., Physicochemical properties of microemulsion analogues<br />

of low density lipoprotein containing amphiphatic apoprotein B receptor sequences, Int.<br />

J. Pharm., 228, 109–117, 2001.<br />

29. Baillie, G., Owens, M. D., and Halbert, G. W., A synthetic low density lipoprotein particle capable of<br />

supporting U937 proliferation in vitro, J. Lipid Res., 43, 69–73, 2002.<br />

30. Markel, A. and Brook, G. J., Cancer and hypocholesterolemia, Isr. J. Med. Sci., 30, 787–793, 1994.<br />

31. Fiorenza, A. M., Branchi, A., and Sommariva, D., Serum lipoprotein profile in patients with cancer.<br />

A comparison with non-cancer subjects, Int. J. Clin. Lab Res., 30, 141–145, 2000.<br />

32. Dessì, S., Batetta, B., Pulisci, D., Accogli, P., Pani, P., and Broccia, G., Total and HDL cholesterol in<br />

human hematologic neoplasms, Int. J. Hematol., 54, 483–486, 1991.<br />

33. Moschovi, M., Trimis, G., Apostolakou, F., Papassotiriou, I., and Tzortzatou-Stathopoulou, F. J.,<br />

Serum lipid alterations in acute lymphoblastic leukemia of childhood, Pediatr. Hematol. Oncol.,<br />

26, 289–293, 2004.<br />

34. Scribano, D., Baroni, S., Pagano, L., Zuppi, C., Leone, G., and Giardina, B., Prognostic relevance of<br />

lipoprotein cholesterol levels in acute lymphocytic and nonlymphocytic leukemia, Haematologica,<br />

81, 343–345, 1996.<br />

35. Siemianowicz, K., Gminski, J., Stajszczyk, M., Wojakowski, W., Goss, M., Machalski, M., Telega,<br />

A., Brulinski, K., and Magiera-Molendowska, H., Serum HDL cholesterol concentration in patients<br />

with squamous cell and small cell lung cancer, Int. J. Mol. Med., 6, 307–311, 2000.<br />

36. Bayerdörffer, E., Mannes, G. A., Richter, W. O., Ochsenkühn, T., Seeholzer, G., Köpcke, W.,<br />

Wiebecke, B., and Paumgartner, G., Decreased high-density lipoprotein cholesterol and increased<br />

low-density cholesterol levels in patients with colorectal adenomas, Ann. Int. Med., 118, 481–487,<br />

1993.<br />

37. Cao, W. M., Murao, K., Imachi, H., Yu, H., Abe, H., Yamauchi, A., Niimi, M., Miyauchi, A., Wong,<br />

N. C., and Ishida, T., A mutant high-density lipoprotein receptor inhibits proliferation of human breast<br />

cancer cells, Cancer Res., 64, 1515–1521, 2004.<br />

38. Lacko, A. G., Nair, M. P., Paranjape, S., Johnson, S., and McConathy, W. J., A novel delivery system<br />

for targeted cancer therapy, In Stem Cell and Targeted Therapy, Dickey, K. A. and Keating, A., Eds.,<br />

Garden Jennings, Charlottesville, VA, pp. 179–183, 2003.<br />

39. Favre, G., Tazi, K. A., Le Gaillard, F., Bennis, F., Hachem, H., and Soula, G., High density lipoprotein3<br />

binding sites are related to DNA biosynthesis in the adenocarcinoma cell line A549, J. Lipid<br />

Res., 34, 1093–1106, 1993.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


784<br />

Nanotechnology for Cancer Therapy<br />

40. Gospodarowicz, D., Lui, G. M., and Gonzalez, R., High-density lipoproteins and the proliferation of<br />

human tumor cells maintained on extracellular matrix-coated dishes and exposed to defined medium,<br />

Cancer Res., 42, 3704–3713, 1982.<br />

41. Jozan, S., Faye, J. C., Tournier, J. F., Tauber, J. P., David, J. F., and Bayard, F., Interaction of estradiol<br />

and high density lipoproteins on proliferation of the human breast cancer cell line MCF-7 adapted to<br />

grow in serum free conditions, Biochem. Biophys. Res. Commun., 133, 105–112, 1985.<br />

42. Pussinen, P. J., Karten, B., Wintersperger, A., Reicher, H., McLean, M., Malle, E., and Sattler, W.,<br />

The human breast carcinoma cell line HBL-100 acquires exogenous cholesterol from high-density<br />

lipoprotein via CLA-1 (CD-36 and LIMPII analogous 1)-mediated selective cholesteryl ester uptake,<br />

Biochem. J., 349, 559–566, 2000.<br />

43. Lacko, A. G., Nair, M., Paranjape, S., Johnson, S., and McConathy, W. J., High density lipoprotein<br />

complexes as delivery vehicles for anticancer drugs, Anticancer Res., 22, 2045–2049, 2002.<br />

44. De Smidt, P. C. and van Berkel, T. J., LDL-mediated drug targeting, Crit. Rev. Ther. Drug Carrier.<br />

Syst., 7, 99–120, 1990.<br />

45. Shaw, J. M. and Shaw, K. V., Key issues in the delivery of pharmacological agents using lipoproteins:<br />

Design of a synthetic apoprotein-lipid carrier, Targeted Diagn. Ther., 5, 351–383, 1991.<br />

46. Barter, P. J., Hugh sinclair lecture: The regulation and remodelling of HDL <strong>by</strong> plasma factors,<br />

Atheroscler. Suppl., 3, 39–47, 2002.<br />

47. Linsel-Nitschke, P. and Tall, A. R., HDL as a target in the treatment of atherosclerotic cardiovascular<br />

disease, Nat. Rev. Drug Discov., 4, 193–205, 2005.<br />

48. Hamilton, R. L., Williams, M. C., Fielding, C. J., and Havel, R. J., Discoidal bilayer structure of<br />

nascent high density lipoproteins from perfused rat liver, J. Clin. Invest., 58, 667–680, 1976.<br />

49. Forte, T., Norum, K. R., Glomset, J. A., and Nichols, A. V., Plasma lipoproteins in familial lecithin:<br />

Cholesterol acyltransferase deficiency: Structure of low and high density lipoproteins as revealed <strong>by</strong><br />

electron microscopy, J. Clin. Invest., 50, 1141–1148, 1971.<br />

50. Lerch, P. G., Fortsch, V., Hodler, G., and Bolli, R., Production and characterization of a reconstituted<br />

high density lipoprotein for therapeutic applications, Vox Sang., 71, 155–164, 1996.<br />

51. Levine, D. M., Simon, S. R., Gordon, B. R., Parker, T. S., Saal, S. D., and Rubin, A. L., United States<br />

Patent no. 5,128,318.<br />

52. Sparks, D. L., Phillips, M. C., and Lund-Katz, S., The conformation of apolipoprotein A-I in discoidal<br />

and spherical recombinant high density lipoprotein particles 13C. NMR studies of lysine ionization<br />

behavior, J. Biol. Chem., 267, 25830–25838, 1992.<br />

53. Jonas, A., Wald, J. H., Toohill, K. L., Krul, E. S., and Kézdy, K. E., The conformation of apolipoprotein<br />

A-I in discoidal and spherical recombinant high density lipoprotein particles. 13C NMR<br />

studies of lysine ionization behavior, J. Biol. Chem., 265, 22123–22129, 1990.<br />

54. Schouten, D., van der Kooij, M., Muller, J., Pieters, M. N., Bijsterbosch, M. K., and van Berkel, T. J.,<br />

Development of lipoprotein-like lipid particles for drug targeting: Neo-high density lipoproteins, Mol.<br />

Pharmacol., 44, 486–492, 1993.<br />

55. Lacko, A. G., Nair, M., Paranjape, S., Mooberry, L., and McConathy, W. J., Advanced Drug Delivery<br />

System for Breast Cancer Chemotherapy Department of Defense, Congressionally Directed Medical<br />

Research Program “ERA of Hope” Breast Cancer Research Review, Philadelphia, Pennsylvania,<br />

2005. http://mrmcweb4.detrick.army.mil/bcrp/era/abstracts2005/0110582_abs.pdf<br />

56. Shaw, J. M., Shaw, K. V., Yanovich, S., Iwanik, M., Futch, W. S., Rosowsky, A., and Schook, L. B.,<br />

Delivery of lipophilic drugs using lipoproteins, Ann. NY Acad. Sci., 507, 252–271, 1987.<br />

57. Adams, T., Alanazi, F., and Lu, D. R., Safety and utilization of blood components as therapeutic<br />

delivery systems, Curr. Pharm. Biotechnol., 4, 275–282, 2003.<br />

58. Nanjee, M. N., Crouse, J. R., King, J. M., Hovorka, R., Rees, S. E., Carson, E. R., Morgenthaler, J. J.,<br />

Lerch, P., and Miller, N. E., Effects of intravenous infusion of lipid-free apo A-I in humans, Arterioscler.<br />

Thromb. Vasc. Biol., 16, 1203–1214, 1996.<br />

59. Nanjee, M. N., Doran, J. E., Lerch, P. G., and Miller, N. E., Acute effects of intravenous infusion of<br />

ApoA1/phosphatidylcholine discs on plasma lipoproteins in humans, Arterioscler. Thromb. Vasc.<br />

Biol., 19, 979–989, 1999.<br />

60. Bisoendial, R. J., Hovingh, G. K., Levels, J. H., Lerch, P. G., Andresen, I., Hayden, M. R., Kastelein,<br />

J. J., and Stroes, E. S., Restoration of endothelial function <strong>by</strong> increasing high-density lipoprotein in<br />

subjects with isolated low high-density lipoprotein, Circulation, 107, 2944–2948, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Lipoprotein Nanoparticles as Delivery Vehicles for Anti-Cancer Agents 785<br />

61. Pajkrt, D., Doran, J. E., Koster, F., Lerch, P. G., Arnet, B., van der Poll, T., ten Cate, J. W., and van<br />

Deventer, S. J., Antiinflammatory effects of reconstituted high-density lipoprotein during human<br />

endotoxemia, J. Exp. Med., 184, 1601–1608, 1996.<br />

62. Nanjee, M. N., Cooke, C. J., Garvin, R., Semeria, F., Lewis, G., Olszewski, W. L., and Miller, N. E.,<br />

J. Lipid Res., 42, 1586–1593, 2001.<br />

63. Palatini, P., Disposition kinetics of phospholipid liposomes, Adv. Exp. Med. Biol., 318, 375–391, 1992.<br />

64. Edwards, I. J., Berquin, I. M., Sun, H., O’Flaherty, J. T., Daniel, L. W., Thomas, M. J., Rudel, L. L.,<br />

Wykle, R. L., and Chen, Y. Q., Differential effects of delivery of omega-3 fatty acids to human cancer<br />

cells <strong>by</strong> low-density lipoproteins versus albumin, Clin. Cancer Res., 10, 8275–8283, 2004.<br />

65. Toi, M., Takada, M., Bando, H., Toyama, K., Yamashiro, H., Horiguchi, S., and Saji, S., Current<br />

status of antibody therapy for breast cancer, Breast Cancer, 11, 10–14, 2004.<br />

66. Boskovitz, A., Wikstrand, C. J., Kuan, C. T., Zalutsky, M. R., Reardon, D. A., and Bigner, D. D.,<br />

Monoclonal antibodies for brain tumour treatment, Expert. Opin. Biol. Ther., 4, 1453–1471, 2004.<br />

67. Bijsterbosch, M. K. and Van Berkel, T. J., Uptake of lactosylated low-density lipoprotein <strong>by</strong> galactose-specific<br />

receptors in rat liver, Biochem. J., 270, 233–239, 1990.<br />

68. Bijsterbosch, M. K. and Van Berkel, T. J., Lactosylated high density lipoprotein: A potential carrier<br />

for the site-specific delivery of drugs to parenchymal liver cells, Mol. Pharmacol., 41, 404–411, 1992.<br />

69. de Vrueh, R. L., Rump, E. T., Sliedregt, L. A., Biessen, E. A., van Berkel, T. T., and Bijsterbosch,<br />

M. K., Synthesis of a lipophilic prodrug of 9-(2-phosphonyl methoxyethyl)adenine (PMEA) and its<br />

incorporation into a hepatocyte-specific lipidic carrier, Pharm. Res., 16, 1179–1185, 1999.<br />

70. de Vrueh, R. L., Rump, E. T., van De Bilt, E., van Veghel, R., Balzarini, J., Biessen, E. A., van Berkel,<br />

T. J., and Bijsterbosch, M. K., Carrier-mediated delivery of 9-(2-phosphonylmethoxyethyl)adenine to<br />

parenchymal liver cells: A novel therapeutic approach for hepatitis B, Antimicrob. Agents<br />

Chemother., 44, 477–483, 2000.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


38<br />

CONTENTS<br />

DQAsomes as Mitochondria-<br />

Targeted Nanocarriers for<br />

Anti-Cancer Drugs<br />

Shing-Ming Cheng, Sarathi V. Boddapati, Gerard<br />

G. M. D’Souza, and Volkmar Weissig<br />

38.1 Introduction ....................................................................................................................... 787<br />

38.2 Apoptosis and Mitochondria ............................................................................................. 788<br />

38.3 Proapoptotic Drugs Acting on Mitochondria ................................................................... 788<br />

38.4 Mitochondriotropic Vesicles (DQAsomes)....................................................................... 790<br />

38.5 DQAsome-Mediated Delivery of pDNA to Mitochondria in Living<br />

Mammalian Cells .............................................................................................................. 791<br />

38.6 Encapsulation of Paclitaxel into DQAsomes.................................................................... 793<br />

38.7 DQAsomal-Encapsulated Paclitaxel Triggers Apoptosis In Vitro ................................... 794<br />

38.8 Tumor Growth Inhibition Study with Paclitaxel-Loaded DQAsomes In Vivo ............... 796<br />

38.9 Concluding Remarks ......................................................................................................... 797<br />

References..................................................................................................................................... 797<br />

38.1 INTRODUCTION<br />

A major challenge in treating cancer is the lack of selectivity of the currently available cytotoxic<br />

agents. Many such agents fail to significantly distinguish between cancer cells and healthy cells.<br />

Consequently, systemic application of these drugs is prone to cause severe side effects in other<br />

tissues, greatly limiting the maximal allowable dose of the drug. Another major hurdle to successfully<br />

treating cancer is multi-drug resistance (MDR). There are multiple putative origins of MDR,<br />

but one of the most studied today is the efflux pump that serves to remove drug molecules from the<br />

cell before they can act at their particular subcellular target inside the cell. Both problems, i.e.,<br />

non-specificity and MDR, could potentially be solved <strong>by</strong> the development of highly selective<br />

tumor-specific drug delivery systems. Progress in this direction has already been made with the<br />

development and subsequent Food and Drug Administration (FDA) approval of Doxil w , the first<br />

liposomal anti-cancer drug, and similar systems are under development. However, the DQAsomebased<br />

approach described in this chapter takes drug delivery one significant step farther. To<br />

efficiently and selectively eradicate carcinoma cells, the cytotoxic drug not only needs to be<br />

delivered to the tumor cell, but it must also be delivered to the particular target inside the cell.<br />

Tumor-specific subcellular drug delivery constitutes a new approach for the chemotherapy<br />

of cancer. This approach becomes increasingly feasible as the particular mechanism of action,<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

787


788<br />

i.e., the molecular targets of anti-cancer drugs, is being recognized. Transporting the cytotoxic drug<br />

to its intracellular target could potentially overcome MDR <strong>by</strong> <strong>by</strong>passing the p-glycoprotein, i.e., the<br />

drug would literally be hidden from the p-glycoprotein inside the delivery system until it becomes<br />

selectively released at the particular intracellular site of action. At the same time, a sub-cellular<br />

delivery system would significantly increase the subcellular bioavailability of any drug acting<br />

inside a cell.<br />

38.2 APOPTOSIS AND MITOCHONDRIA<br />

Apoptosis (programmed cell death) plays a central role in tissue homeostasis, and it is generally<br />

recognized that inhibition of apoptosis may contribute to cell transformation. 1 Significant knowledge<br />

has been accumulated during the past several years about how the apoptotic machinery is<br />

controlled. 2–20 As each new regulatory mechanism had been identified, dysfunction of that<br />

mechanism has been linked to one or another type of cancer. 17 Dysregulation of the apoptotic<br />

machinery is now generally accepted as an almost universal component of the transformation<br />

process of normal cells into cancer cells.<br />

A large body of experimental data demonstrates that mitochondria play a key role in the<br />

complex apoptotic mechanism. 18–49 Mitochondria have been shown to trigger cell death via<br />

several mechanisms: <strong>by</strong> disrupting electron transport and energy metabolism, <strong>by</strong> releasing or<br />

activating proteins that mediate apoptosis, and <strong>by</strong> altering cellular redox potential. A critical<br />

event leading to programmed cell death is the mitochondrial membrane permeabilization that is<br />

under the control of the permeability transition pore complex (mPTPC), a multiprotein complex<br />

formed at the contact site between the mitochondrial inner and outer membranes. The mPTPC is<br />

widely accepted as being central to the process of cell death and has accordingly been recommended<br />

as a privileged pharmacological target for cytoprotective and for cytotoxic therapies<br />

in general. 50 In particular, it has been suggested that targeting specific mPTPC components may<br />

overcome bcl-2 mediated apoptosis inhibition in cancer cells. 1 Several studies have already demonstrated<br />

the feasibility of eliminating neoplastic cells <strong>by</strong> selectively inducing apoptosis (reviewed in<br />

reference 17). The design of mitochondria-targeted cytotoxic drugs has been formulated as a novel<br />

strategy for overcoming apoptosis resistance in tumor cells, 1 which, intriguingly, opens up a whole<br />

new avenue for the therapy of cancer <strong>by</strong> “tricking cancer cells into committing suicide.” 17<br />

38.3 PROAPOPTOTIC DRUGS ACTING ON MITOCHONDRIA<br />

Nanotechnology for Cancer Therapy<br />

Several conventional anti-cancer drugs, such as doxorubicin, and cisplatin, have no direct effect on<br />

mitochondria. 51 These conventional chemotherapeutic agents elicit mitochondrial permeabilization<br />

in an indirect fashion <strong>by</strong> induction of endogenous effectors that are involved in the physiologic<br />

control of apoptosis. 1 However, a variety of clinically approved drugs such as paclitaxel, 52–60<br />

VP-16 (etoposide) 61–64 and vinorelbine 58 as well as an increasing number of experimental anticancer<br />

drugs such as betulinic acid, lonidamine, CD-437 (a synthetic retinoids) and ceramide<br />

(reviewed in 1 ) have been found to act directly on mitochondria resulting in triggering apoptosis.<br />

These agents may induce apoptosis in circumstances in which conventional drugs fail to act<br />

because endogenous apoptosis inducing pathways, e.g., such as those involving p53, death<br />

receptors or apical caspase activation, are disrupted, leading to the apoptosis-resistance of tumor<br />

cells. For example, several in vitro and in vivo studies have shown that the synthetic retinoid CD437<br />

is able to induce apoptosis in human lung, breast, cervical and ovarian carcinoma cells (reviewed in<br />

Kaufmann and Gores 2000). It could be demonstrated that in intact cells, CD437-dependent caspase<br />

activation is preceded <strong>by</strong> the release of cytochrome C from mitochondria. 65 Moreover, it was<br />

shown that when added to isolated mitochondria, CD437 causes membrane permeabilization and<br />

that this effect is prevented <strong>by</strong> inhibitors of the mPTPC such as cyclosporine A. CD437 constitutes<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 789<br />

an experimental drug that exerts its cytotoxic effect via the mPTPC, i.e., <strong>by</strong> acting directly at the<br />

surface or inside of mitochondria.<br />

The development of anti-cancer drugs whose cytotoxic effects depend on their direct interaction<br />

with mitochondria inside of living cells raises the issue of the intracellular distribution of these<br />

drugs after having been taken up <strong>by</strong> the cell, i.e., the question of their intracellular bioavailability.<br />

Independent of their mode of cell entry that should mostly take place via passive diffusion through<br />

the cell membrane, drug molecules become randomly distributed among all cell organelles and,<br />

depending on the chemical nature of the drug, eventually metabolized. Therefore, anti-cancer drugs<br />

that exert their cytotoxic activity <strong>by</strong> directly acting at or inside of mitochondria in living cells would<br />

dramatically benefit from a delivery system that can selectively transport these drugs to and<br />

into mitochondria.<br />

One of the major roles of mitochondria in the metabolism of eukaryotic cells is the synthesis of<br />

Adenosine triphosphate (ATP) <strong>by</strong> oxidative phosphorylation via the respiratory chain. According to<br />

Mitchell’s chemiosmotic hypothesis, electrons from the hydrogens on Nicotine amide adenine<br />

dinucleotide (NADH) and Flavin adenine dinucleotide (FADH 2) are carried along the respiratory<br />

chain at the mitochondrial inner membrane, there<strong>by</strong> releasing energy that is used to pump protons<br />

across the inner membrane from the mitochondrial matrix into the intermembrane space. This<br />

process creates a transmembrane electrochemical gradient that includes contributions from both<br />

a membrane potential (negative inside) and a pH difference (acidic outside). The membrane<br />

potential of mitochondria in vitro is between 180 and 200 mV, the maximum a lipid bilayer can<br />

sustain while maintaining its integrity. 66 Although this potential is reduced in living cells and<br />

organism to about 130–150 mV as a result of metabolic processes such as ATP synthesis and<br />

ion transport, 67 it is <strong>by</strong> far the largest within cells.<br />

Most interestingly, carcinoma cells posses a different mitochondrial membrane potential<br />

relative to normal cells. It has been found that in many carcinoma cell lines, the mitochondrial<br />

membrane potential is higher than in normal epithelial cells (reviewed in reference 68). For<br />

example, the difference of the mitochondrial membrane potential between the colon carcinoma<br />

cell line CX-1 and the control green monkey kidney epithelial cell line CV-1 has been reported to<br />

be approximately 60 mV. 68 Moreover, some carcinoma cells, in particular, human breast adenocarcinoma-derived<br />

cells, have in addition to the higher mitochondrial membrane potential also an<br />

elevated plasma membrane potential relative to their normal parent cell lines. 68–75<br />

The striking difference between normal cells and human adenocarcinoma cells regarding the<br />

electrical charge of both plasma and mitochondrial membranes has lead numerous investigators<br />

during the 1990s to explore fundamentally new strategies for the selective targeting of cancer cells.<br />

Their attempts have been based on compounds with a delocalized charge that have long been<br />

known to accumulate in mitochondria of living cells in response to the mitochondrial membrane<br />

potential. Many of these DLCs are toxic to mitochondria at high concentration. For example,<br />

the rhodacyanine official name of this compound, unknown what the letters stand for<br />

(MKT-077 71,76–80 ) was the first DLC to be approved <strong>by</strong> the FDA for clinical trials for the treatment<br />

of carcinoma. 81 The trials were discontinued, however, because efficacy in tumor cell killing was<br />

not demonstrated at the particular approved dosage and drug regimen. 68 From the phase I clinical<br />

trial, it was concluded that it is feasible to target carcinoma cell mitochondria with rhodacyanine<br />

analogues if drugs with higher therapeutic indices could be developed. 81<br />

The DQAsomal-based strategy involving the use of dequalinium chloride, a typical representative<br />

of DLCs, for tumor and mitochondria-specific targeting factually starts where these failed<br />

attempts of the 1990s have left off. This approach combines the well-proven ability of DLCs to<br />

specifically target carcinoma cell mitochondria with the selective delivery of apoptotically active<br />

compounds known to trigger programmed cell death via directly acting at the mitochondrial<br />

surface. The DQAsomal approach utilizes the mitochondria-specific affinity of DLCs for the<br />

delivery of pro-apoptotic drugs to mitochondria. Therefore, the actual cytotoxic effect is caused<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


790<br />

<strong>by</strong> apoptosis-triggering drugs. Any inherent cytotoxicity of the carrier system, i.e., the DLCs used,<br />

might only add to the overall efficiency of killing carcinoma cells.<br />

38.4 MITOCHONDRIOTROPIC VESICLES (DQASOMES)<br />

Dequalinium chloride (DQA, Figure 38.1a) represents a typical mitochondriotropic delocalized<br />

cation that already almost 20 years ago was shown to selectively accumulate in carcinoma cell<br />

mitochondria. 73 Most interestingly in the context of this chapter, it was demonstrated recently that<br />

dequalinium B, a new boron carrier for neutron capture therapy, also accumulates preferentially in<br />

carcinoma cells over non-transformed cells. 82<br />

Dequalinium is a dicationic compound resembling “bola”-form electrolytes, i.e., it is a symmetrical<br />

molecule with two charge centers separated at a relatively large distance. Such symmetric<br />

bola-like structures are well known from archaeal lipids that usually consist of two glycerol<br />

backbones connected <strong>by</strong> two hydrophobic chains. 85,86 The self-assembly behavior of bipolar<br />

lipid from Archaea has been extensively studied (reviewed in Gambacorta, Gliozi, and De Rosa<br />

1995). Generally, it has been shown that these symmetric bipolar archaeal lipids can self-associate<br />

into mechanically, very stable monolayer membranes. The most striking structural difference<br />

between dequalinium and archaeal lipids lies in the number of bridging hydrophobic chains<br />

between the polar head groups. Contrary to the common arachaeal lipids, in dequalinium there<br />

is only one carbohydrate chain that connects the two cationic hydrophilic head groups. Therefore,<br />

this type of bola lipids has been named single-chain bola-amphiphile. 87,88 The self-association<br />

behavior of this single-chain cationic bola amphiphile was investigated using Monte Carlo<br />

computer simulations (Figure 38.1c, left panel), 83 several electron microscopic (EM) techniques<br />

(Figure 38.1c, right panel) as well as dynamic laser light scattering. 84<br />

It was found that, upon sonication, dequalinium forms spherical aggregates with diameters<br />

between about 70 and 700 nm that were termed DQAsomes. 84 Freeze fracture images<br />

(Figure 38.1) show both convex and concave fracture faces. These images strongly indicate the<br />

(a)<br />

(b)<br />

(c)<br />

NH 2<br />

CH 3<br />

0.2μm<br />

OR<br />

CH 3<br />

N CH CH CH CH CH CH CH CH CH 2 2 2 2 2 2 2 2 CH N<br />

2<br />

Nanotechnology for Cancer Therapy<br />

NH 2<br />

A B C<br />

FIGURE 38.1 (a) Chemical structure of dequalinium chloride with overlaid colors indicating, in blue, the<br />

hydrophilic part and in yellow the hydrophobic part of the molecule. (b) Theoretical possible conformations of<br />

dequalinium chloride, i.e., stretched versus horseshoe conformation, leading to either a monolayer or a bilayer<br />

membranous structure following the process of self-assembly. (c) Left panel: Monte Carlo Computer<br />

Simulations demonstrate the possible self-assembly of dequalinium chloride into vesicles. (From Weissig,<br />

V., Mogel, H. J., Wahab, M., and Lasch, J., Proceed. Intl. Symp. Control. Rel. Bioact. Mater., 25, 312, 1998.)<br />

Right panel: Electron microscopic images of vesicles (DQAsomes) made from dequalinium chloride, from left<br />

to right: Negatively stained, rotary shadowed, freeze fractured. (From Weissig, V., Lasch, L., Erdos, G.,<br />

Meyer, H. W., Rowe, T. C., and Hughes, J., Pharm. Res., 15(2), 334–337, 1998. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 791<br />

FIGURE 38.2 Top panel: Structure of the cyclohexyl derivative of dequalinium. Bottom panel: Schematic<br />

illustration of the stabilizing effect of the cyclohexyl ring system (black circles).<br />

liposome-like aggregation of dequalinium. Negatively stained samples (Figure 38.1) demonstrate<br />

that the vesicle is impervious to the stain and appears as a clear area surrounded <strong>by</strong> stain with no<br />

substructure visible. Particle size measurements of DQAsomes stored at room temperature for 24 and<br />

96 h (not shown) do not show any significant changes in their size distribution in comparison to<br />

freshly made vesicles measured after one hour. 84 This indicates that DQAsomes do not seem to<br />

precipitate, to fuse with each other, or to aggregate in solution over a period of several days. The use<br />

of dequalinium derivatives for the preparation of DQAsome-like vesicles leads to vesicles with<br />

different size distributions. 89 For example, substituting the methyl group <strong>by</strong> an aliphatic ring<br />

system (Figure 38.2) confers superior vesicle forming properties to this bolaamphiphile. These<br />

DQAsome-like vesicles, i.e., vesicles prepared from the cyclohexyl-derivative of DQA shown in<br />

Figure 38.2, have a very narrow size distribution of 169G50 nm and can be stored at room temperature<br />

for at least five months.<br />

In contrast to vesicles made from dequalinium, bolasomes made from the cyclohexyl derivative<br />

are also stable upon dilution of the original vesicle preparation. Whereas dequalinium-based bolasomes<br />

upon dilution slowly disintegrate over a period of several hours, bolasomes made from the<br />

cyclohexyl compound do not show any change in size distribution following dilution. It appears<br />

that bulky aliphatic residues attached to the quinolinium heterocycle favor self-association of the<br />

planar ring system. It has, therefore, been speculated that the bulky group sterically prevents the<br />

free rotation of the hydrophilic head of the amphiphile around the CH2-axis (Figure 38.2, bottom<br />

panel), contributing to improved intermolecular interactions between the amphiphilic monomers.<br />

38.5 DQASOME-MEDIATED DELIVERY OF pDNA TO MITOCHONDRIA<br />

IN LIVING MAMMALIAN CELLS<br />

During efforts in developing a mitochondria-specific DNA delivery vector, it could be demonstrated<br />

that DQAsomes are able to selectively deliver plasmid DNA (pDNA) to mitochondria<br />

within living mammalian cells. It has been shown, in particular, that DQAsomes stably incorporate<br />

pDNA, 84 protect the pDNA from nuclease digestion, and mediate its cellular uptake most likely via<br />

non-specific endocytosis. 90 Using membrane-mimicking liposomal membranes and isolated rat<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


792<br />

Nanotechnology for Cancer Therapy<br />

liver mitochondria, it was shown that DQAsome/pDNA complexes become destabilized upon<br />

contact with mitochondrial membranes, but not at cell plasma membranes. 91,92<br />

The selective destabilization of DQAsomes at mitochondrial membranes may either be caused<br />

<strong>by</strong> the difference in the lipid composition between cytoplasmic and mitochondrial membranes 91 or<br />

<strong>by</strong> the membrane potential-driven diffusion of individual dequalinium molecules from the<br />

DQAsome/pDNA complex into the mitochondrial matrix leading to the disintegration of the<br />

complex. However, the fact that DQAsomes do not lose their cargo (pDNA) during their contact<br />

with the plasma cell membrane (i.e., during their cellular uptake) but do release the entrapped<br />

pDNA upon contact with mitochondria appears as a very attractive feature of DQAsomes as a<br />

mitochondria-specific drug delivery system. Any encapsulated drug would potentially be released<br />

upon contact with mitochondrial membranes leading to the desired high local drug concentration in<br />

the immediate proximity to the mPTPC that is recognized as a major target for apoptotically active<br />

experimental drugs. 1,93,94<br />

Before translocating to mitochondria in response to the mitochondrial membrane potential,<br />

however, endocytosed DQAsomes have to be released from endosomes into the cytosol. From<br />

studies about the intracellular fate of cationic liposome/DNA complexes (lipoplexes), it is known<br />

that cationic lipids exert a destabilizing effect on endosomal membranes, leading to the release of at<br />

least a fraction of the lipoplex from early endosomes. 95–98 In agreement with these data, it was<br />

found that dequalinium-based DQAsomes also display endosomolytic activity. 99 It was shown that<br />

adding DQAsomes to liposomes mimicking the lipid composition of endosomal membranes reproducibly<br />

lead to the release of liposomal encapsulated fluorescence marker. 99<br />

Direct evidence for the ability of DQAsomes to transport pDNA selectively to the site of<br />

mitochondria was provided <strong>by</strong> studying the intracellular distribution of mitochondrial leader<br />

sequence peptide -pDNA conjugates in cultured BT20 cells using confocal fluorescence<br />

microscopy. 100 Figure 38.3 shows images representative of the confocal fluorescence micrographs<br />

obtained in this study. The green and red channels used to generate the overlaid images in the far<br />

right column are shown separately in the preceding columns to facilitate a careful comparison of the<br />

observed staining patterns.<br />

The characteristic mitochondrial staining pattern seen with the red channel is a strong indicator<br />

of mitochondrial viability in the imaged cells. From the composite image obtained <strong>by</strong> overlaying<br />

the green and red channels, it can be seen that a sizeable fraction of the intracellular green<br />

fluorescence co-localized with the red mitochondrial fluorescence (depicted as white areas in<br />

Figure 38.3c). These observations indicate that, in addition to mediating the cellular uptake of<br />

the pDNA conjugate, the use of DQAsomes resulted in a definite association of an appreciable<br />

fraction of the internalized conjugate with mitochondria.<br />

FIGURE 38.3 (See color insert following page 522.) Representative confocal fluorescence micrographs of<br />

BT20 cells stained with Mitotrackerw Red CMXRos (red) after exposure to fluorescein labeled linearized<br />

MLS-pDNA conjugate (green) complexed with DQAsomes; (a) red channel, (b) green channel, (c) overlay of<br />

red and green channels with white, indicating co-localization of red and green fluorescence.(From D’Souza, G.<br />

G., Boddapati, S. V., and Weissig, V., Mitochondrion, 5(5), 352–358, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 793<br />

38.6 ENCAPSULATION OF PACLITAXEL INTO DQASOMES<br />

Paclitaxel, generally known as an anti-microtubule agent, has recently been demonstrated to trigger<br />

apoptosis <strong>by</strong> directly acting on mitochondria. 54,55,58 It has been shown that clinically relevant<br />

concentrations of paclitaxel directly target mitochondria and trigger apoptosis <strong>by</strong> inducing<br />

cytochrome c (cyt c) release in a permeability transition pore (PTP)-dependent manner. 54 This<br />

mechanism of action is known from other pro-apoptotic, directly on mitochondria acting agents. 51<br />

A 24-hour delay between the treatment with paclitaxel or with other PTP inducers and the release of<br />

cyt c in cell-free systems compared to intact cells has been explained <strong>by</strong> the existence of several<br />

drug targets inside the cell. Making only a subset of the drug available for mitochondria. 54 Likewise,<br />

other anti-tubulin agents such as vinorelbine or nocodazole could also been shown to trigger<br />

the release of cyt c via the direct interaction with mitochondria 101 that subsequently resulted in<br />

apoptotic cell death. The detection of tubulin as an inherent component of mitochondrial<br />

membranes able to interact with the voltage-dependent anion channel 102 and there<strong>by</strong> able to<br />

influence the mPTPC seems to make the rather surprising identification of mitochondria as a<br />

target for well-established anti-tubulin agents plausible.<br />

For encapsulation of paclitaxel into DQAsomes, dequalinium chloride and paclitaxel were<br />

dissolved in methanol followed <strong>by</strong> removing the organic solvent. 103 After adding buffer, the<br />

suspension was sonicated with a probe sonicator until a clear opaque solution was formed. To<br />

remove any undissolved material, the sample was centrifuged for 10 min at 3000 rpm. The solubility<br />

of paclitaxel in water at 258C at pH 7.4 is with 0.172 mg/L (0.2 mM) extremely low, making<br />

any separation procedure of non-encapsulated paclitaxel unnecessary. However, for control, a<br />

paclitaxel suspension was probe sonicated under identical conditions used for the incorporation<br />

of paclitaxel into DQAsomes but in the complete absence of dequalinium. As expected, upon<br />

centrifugation, no paclitaxel was detectable in the supernatant using UV spectroscopy at<br />

230 nm. 103 Following this procedure, paclitaxel can be incorporated into DQAsomes at a molar<br />

ratio paclitaxel to dequalinium of about 0.6. In comparison to the free drug, encapsulation of<br />

paclitaxel into DQAsomes increases the drug’s solubility <strong>by</strong> a factor of about 3000.<br />

Weight distribution analysis<br />

90%<br />

50%<br />

10%<br />

Dust<br />

0.00%<br />

200 600 1000<br />

Size [nm]<br />

FIGURE 38.4 Paclitaxel encapsulated into DQAsomes. Left panel: transmission electron microscopic image<br />

(uranyl acetate staining); middle panel: Size distribution; right panel: Cryo-electron microscopic image. (From<br />

Cheng, S. M., Pabba, S., Torchilin, V. P., Fowle, W., Kimpfler, A., Schubert, R. and Weissig, V. J. Drug Deliv.<br />

Sci. Technol., 15(1), 81–86, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


794<br />

Paclitaxel / DQA (molar)<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

Day 1 Day 3<br />

Day 4 Day 6 Day 10<br />

Nanotechnology for Cancer Therapy<br />

FIGURE 38.5 Paclitaxel remains stably incorporated in DQAsomes: Molar ratio of paclitaxel to DQA in<br />

DQAsomes upon storage at 4C. (From Cheng, S. M., Pabba, S., Torchilin, V. P., Fowle, W. Kimpfler, A.<br />

Schubert, R., and Weissig, V., J. Drug Deliv. Sci. Technol., 15(1), 81–86, 2005. With permission.)<br />

Considering the known spherical character of DQAsomes, the results of an EM analysis of<br />

paclitaxel-loaded DQAsomes seem rather surprising. The transmission EM image (Figure 38.4, left<br />

panel) and the cryo-EM image (Figure 38.4, right panel) of an identical sample show with a<br />

remarkable conformity worm- or rod-like structures roughly around 400 nm in length, the size<br />

of which could also be confirmed <strong>by</strong> size distribution analysis shown in Figure 38.4 (middle panel).<br />

These complexes may represent the formation of worm-like micelles as recently described for selfassembling<br />

amphiphilic block co-polymers. 104 Figure 38.5 shows data about the stability of<br />

paclitaxel-containing DQAsomes upon storage. Over a period of ten days, the ratio of paclitaxel<br />

to dequalinium in the DQAsome preparation remains almost constant, indicating that paclitaxel<br />

remains stably incorporated in DQAsomes. However, from a gradual decrease of the total concentration<br />

of paclitaxel and of dequalinium over several days and to the same extent (<strong>by</strong> approximately<br />

20–30% after four days, data not shown) can be concluded that paclitaxel-loaded DQAsomes tend<br />

to form larger aggregates that are removable <strong>by</strong> centrifugation. This tendency to form aggregates is<br />

well known from artificial phospholipids vesicles (liposomes) and is generally reversible <strong>by</strong><br />

shaking or slightly vortexing the preparation.<br />

The stability of paclitaxel-loaded DQAsomes is under physiological conditions in comparison<br />

to their shelf-life stability significantly reduced. In general, upon transferring paclitaxel-loaded<br />

DQAsomes either into serum-free cell medium (DMEM) at 378C or into complete serum, they<br />

form large aggregates leading to precipitation. However, this process of aggregation seems to be<br />

kinetically controlled, giving at least a portion of monomeric paclitaxel-loaded DQAsomes sufficient<br />

time to reach their target. The tendency to form aggregates under physiological conditions is<br />

common to all cationic carrier systems. Nevertheless, it should be mentioned that cationic lipoplexes<br />

and polyplexes are being tested in numerous gene therapeutic clinical trials. Also, cationic<br />

colloidal carriers (Catioms) are currently being successfully developed for systemic drug administrations.<br />

105–107<br />

38.7 DQASOMAL-ENCAPSULATED PACLITAXEL TRIGGERS APOPTOSIS<br />

IN VITRO<br />

To test if DQAsomal-encapsulated paclitaxel triggers apoptosis at paclitaxel concentrations where<br />

the free drug does not have a significant cytotoxic effect, human colo 205 colon cancer cells were<br />

incubated with the free drug, with empty DQAsomes, with a mixture of empty DQAsomes and<br />

the free drug, and with the DQAsomal-encapsulated drug. Following the staining of the treated cells<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 795<br />

% of apoptosis cells<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Empty DQAsomes mixture DQAsomes +<br />

free paclitaxel<br />

DQAsomal<br />

encapsulated<br />

paclitaxel<br />

FIGURE 38.6 Human colo 25 colon cancer cells were treated in triplicates with empty DQAsomes (20 nM<br />

DQA) with a mixture of empty DQAsomes (20 nM) and free paclitaxel (10 nM) and DQAsomal-encapsulated<br />

paclitaxel (20 nM DQA/10 nM paclitaxel). In each case, approximately 400 cells were counted (From Chen, S.<br />

M. and Weissig, V., <strong>2006</strong>, manuscript in preparation).<br />

with the DNA-binding fluorophore Hoechst 33258, apoptotic nuclei showing the typical apoptotic<br />

condensation and fragmentation of chromatin were counted and expressed as percent of the total<br />

number of nuclei. Figure 38.6 shows that under identical incubation conditions, 10 nM paclitaxelencapsulated<br />

in DQAsomes more than doubles the number of apoptotic nuclei in comparison to the<br />

control whee cells were treated with a mixture of empty DQAsomes and 10 nM free paclitaxel.<br />

Likewise, a DNA ladder caused <strong>by</strong> DNA fragmentation typical for apoptosis could be detected<br />

upon incubation of colon cancer cells with 10 nM DQAsomal-encapsulated paclitaxel but not upon<br />

incubation with the free drug either alone or in mixture with empty DQAsomes (Figure 38.7).<br />

Incubating the cells for the same period of time, the amount of free paclitaxel had to be increased at<br />

least 5-fold over the amount of DQAsomal-encapsulated drug in order to generate a DNA ladder<br />

(not shown).<br />

FIGURE 38.7 Human colo 25 colon cancer cells were incubated for 30 h with buffer (lane 2), 20 nM empty<br />

DQAsomes (lane 3), 10 nM free paclitaxel (lane 4), a mixture of 20 nM empty DQAsomes and 10 nM free<br />

paclitaxel (lane 5), and 20 nM DQAsomes with 10 nM encapsulated paclitaxel (lane 6). White arrow heads<br />

indicate the apoptotic DNA ladder (Chen, S. M. and Weissig, V., <strong>2006</strong>, manuscript in preparation).<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


796<br />

Considering that paclitaxel, generally known as an anti-microtubule agent, has recently been<br />

demonstrated to trigger apoptosis <strong>by</strong> directly acting on mitochondria, 54,55,58 it can be concluded<br />

from the data shown in Figure 38.6 and Figure 38.7 that encapsulating paclitaxel into a<br />

mitochondria-specific drug delivery system appears to increase the sub-cellular, i.e., mitochondrial,<br />

bioavailability of the drug.<br />

38.8 TUMOR GROWTH INHIBITION STUDY WITH PACLITAXEL-LOADED<br />

DQASOMES IN VIVO<br />

Paclitaxel-loaded DQAsomes were tested for their ability to inhibit the growth of human colon<br />

cancer cells in nude mice. 103 COLO-205 cells were inoculated s.c. into the left flank of nude mice<br />

that all formed palpable tumors within seven days after cell injection. For controls with free<br />

paclitaxel, the drug was suspended in 100% DMSO at 20 mM, stored at 48C, and immediately<br />

before use, diluted in warm medium. In all controls, the dose of free paclitaxel and empty<br />

DQAsomes, respectively, was adjusted according to the dose of paclitaxel and dequalinium<br />

given in the paclitaxel-loaded DQAsome sample. Because of the lack of any inhibitory effect on<br />

tumor growth, the dose was tripled after 1.5 weeks. Figure 38.8 shows that at concentrations where<br />

free paclitaxel and empty DQAsomes do not show any impact on tumor growth, paclitaxel-loaded<br />

DQAsomes (with paclitaxel and dequalinium concentrations identical to controls) seem to inhibit<br />

the tumor growth <strong>by</strong> about 50%. Correspondingly, the average tumor weight in the treatment<br />

group after sacrificing the animals after 26 days is approximately half of that in all controls<br />

(Figure 38.9). Although this result seems to suggest that DQAsomes might be able to increase<br />

the therapeutic potential of paclitaxel, the preliminary character of this first in vivo study should<br />

be emphasized.<br />

Average tumor volume [mm 3 ]<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0 10 20<br />

Day after tumor implantation<br />

30<br />

Nanotechnology for Cancer Therapy<br />

Hepes buffer<br />

Free paclitaxel<br />

Empty DQAsomes<br />

Paclitaxel-loaded<br />

DQAsomes<br />

FIGURE 38.8 Tumor growth inhibition study in nude mice implanted with human colon cancer cells. The<br />

mean tumor volume from each group was blotted against the number of days. Each group involved eight<br />

animals. For clarity, error bars were omitted. Note that after 1.5 weeks, the dose, normalized for paclitaxel, was<br />

tripled in all treatment groups. (From Cheng, S. M., Pabba, S., Torchilin, V. P., Fowle, W., Kimpfler, A.,<br />

Schubert, R., and Weissig, V., J. Drug Deliv. Sci. Technol., 15(1), 81–86, 2005. With permission.)<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 797<br />

38.9 CONCLUDING REMARKS<br />

DQAsomes and DQAsome-like vesicles have been established as the first mitochondria-specific<br />

cationic drug delivery system potentially able to deliver cytotoxic drugs selectively to mitochondria<br />

in cancer cells. However, the encapsulation of antineoplastic drugs into cationic vesicles has<br />

already been suggested in 1998. 108 It was found that upon intravenous injection of gold-labeled<br />

cationic liposomes, 32% associated with tumor endothelial cells, 53% were internalized into<br />

endosomes, and 15% were extravascular 20 min after injection. 108 Following these early experimental<br />

data, Munich Biotech AG (Neuried, Germany) is currently developing so-called Catioms for<br />

the systemic delivery of anti-cancer drugs to solid tumors. 105–107 However, whereas Munich<br />

Biotech is focusing on the targeting of the tumor vasculature, the DQAsomal strategy described<br />

in this chapter is aimed at the delivery of pro-apoptotic compounds to and into tumor<br />

cell mitochondria.<br />

REFERENCES<br />

Average tumor weight [g]<br />

2.000<br />

1.600<br />

1.200<br />

0.800<br />

0.400<br />

0.000<br />

Hepes buffer Free paclitaxel Empty Paclitaxel-Loaded<br />

DQAsomes DQAsomes<br />

FIGURE 38.9 Average tumor weight (nZ8) at time of sacrifice of nude mice implanted with human colon<br />

cancer cells. For treatment group (paclitaxel-loaded DQAsomes) versus all three control groups combined<br />

PZ0.054 (parametric Student t-test). (From Cheng, S. M., Pabba, S., Torchilin, V. P., Fowle, W., Kimpfler,<br />

A., Schubert, R., and Weissig, V., J. Drug Deliv. Sci. Technol., 15(1), 81–86, 2005. With permission.)<br />

1. Costantini, P., Jacotot, E., Decaudin, D., and Kroemer, G., Mitochondrion as a novel target of<br />

anticancer chemotherapy, J. Natl. Cancer Inst., 92(13), 1042–1053, 2000.<br />

2. Galati, G., Teng, S., Moridani, M. Y., Chan, T. S., and O’Brien, P. J., Cancer chemoprevention and<br />

apoptosis mechanisms induced <strong>by</strong> dietary polyphenolics, Drug Metabol. Drug Interact., 17(1–4),<br />

311–349, 2000.<br />

3. Ghafourifar, P., Bringold, U., Klein, S. D., and Richter, C., Mitochondrial nitric oxide synthase,<br />

oxidative stress and apoptosis, Biol. Signals Recept., 10(1–2), 57–65, 2001.<br />

4. Davis, W., Ronai, Z., and Tew, K. D., Cellular thiols and reactive oxygen species in drug-induced<br />

apoptosis, J. Pharmacol. Exp. Ther., 296(1), 1–6, 2001.<br />

5. Creagh, E. M., Sheehan, D., and Cotter, T. G., Heat shock proteins—modulators of apoptosis in<br />

tumour cells, Leukemia, 14(7), 1161–1173, 2000.<br />

6. Hall, A. G., Review: The role of glutathione in the regulation of apoptosis, Eur. J. Clin. Invest.,<br />

29(3), 238–245, 1999.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


798<br />

Nanotechnology for Cancer Therapy<br />

7. Deigner, H. P. and Kinscherf, R., Modulating apoptosis: Current applications and prospects for future<br />

drug development, Curr. Med. Chem., 6(5), 399–414, 1999.<br />

8. Peltenburg, L. T., Radiosensitivity of tumor cells. Oncogenes and apoptosis, Q. J. Nucl. Med., 44(4),<br />

355–364, 2000.<br />

9. Deigner, H. P., Haberkorn, U., and Kinscherf, R., Apoptosis modulators in the therapy of neurodegenerative<br />

diseases, Expert Opin. Investig. Drugs, 9(4), 747–764, 2000.<br />

10. Kanduc, D., Mittelman, A., Serpico, R., Sinigaglia, E., Sinha, A. A., Natale, C., Santacroce, R.,<br />

Di Corcia, M. G. et al., Cell death: Apoptosis versus necrosis (review), Int. J. Oncol., 21(1),<br />

165–170, 2002.<br />

11. Robertson, J. D., Orrenius, S., and Zhivotovsky, B., Review: Nuclear events in apoptosis, J. Struct.<br />

Biol., 129(2,3), 346–358, 2000.<br />

12. Tomei, L. D. and Umansky, S. R., Apoptosis and the heart: A brief review, Ann. N. Y. Acad. Sci., 946,<br />

160–168, 2001.<br />

13. Sathasivam, S., Ince, P. G., and Shaw, P. J., Apoptosis in amyotrophic lateral sclerosis: A review of<br />

the evidence, Neuropathol. Appl. Neurobiol., 27(4), 257–274, 2001.<br />

14. Tinaztepe, K., Ozen, S., Gucer, S., and Ozdamar, S., Apoptosis in renal disease: A brief review of the<br />

literature and report of preliminary findings in childhood lupus nephritis, Turk. J. Pediatr., 43(2),<br />

133–138, 2001.<br />

15. Zusman, I., Gurevich, P., Gurevich, E., and Ben-Hur, H., The immune system, apoptosis and<br />

apoptosis-related proteins in human ovarian tumors (a review), Int. J. Oncol., 18(5), 965–972, 2001.<br />

16. Belzacq, A. S., Vieira, H. L., Kroemer, G., and Brenner, C., The adenine nucleotide translocator in<br />

apoptosis, Biochimie, 84(2,3), 167–176, 2002.<br />

17. Kaufmann, S. H. and Gores, G. J., Apoptosis in cancer: Cause and cure, Bioessays, 22(11),<br />

1007–1017, 2000.<br />

18. Olson, M. and Kornbluth, S., Mitochondria in apoptosis and human disease, Curr. Mol. Med., 1(1),<br />

91–122, 2001.<br />

19. Lin, Q. S., Mitochondria and apoptosis, Sheng Wu Hua Xue Yu Sheng Wu Wu Li Xue Bao (Shanghai),<br />

31(2), 116–118, 1999.<br />

20. Pollack, M. and Leeuwenburgh, C., Apoptosis and aging: Role of the mitochondria, J. Gerontol. A<br />

Biol. Sci. Med. Sci., 56(11), B475–B482, 2001.<br />

21. Henkart, P. A. and Grinstein, S., Apoptosis: Mitochondria resurrected? J. Exp. Med., 183(4),<br />

1293–1295, 1996.<br />

22. Skulachev, V. P., Why are mitochondria involved in apoptosis? Permeability transition pores and<br />

apoptosis as selective mechanisms to eliminate superoxide-producing mitochondria and cell, FEBS<br />

Lett., 397(1), 7–10, 1996.<br />

23. Petit, P. X., Zamzami, N., Vayssiere, J. L., Mignotte, B., Kroemer, G., and Castedo, M., Implication<br />

of mitochondria in apoptosis, Mol. Cell. Biochem., 174(1,2), 185–188, 1997.<br />

24. Decaudin, D., Marzo, I, Brenner, C., and Kroemer, G., Mitochondria in chemotherapy-induced<br />

apoptosis: A prospective novel target of cancer therapy (review), Int. J. Oncol., 12(1), 141–152,<br />

1998.<br />

25. Mignotte, B. and Vayssiere, J. L., Mitochondria and apoptosis, Eur. J. Biochem., 252(1), 1–15, 1998.<br />

26. Susin, S. A., Zamzami, N., and Kroemer, G., Mitochondria as regulators of apoptosis: Doubt no<br />

more, Biochim. Biophys. Acta, 1366(1,2), 151–165, 1998.<br />

27. Green, D. R. and Reed, J. C., Mitochondria and apoptosis, Science, 281(5381), 1309–1312, 1998.<br />

28. Tatton, W. G. and Chalmers-Redman, R. M., Mitochondria in neurodegenerative apoptosis: An<br />

opportunity for therapy? Ann. Neurol., 44(3 Suppl. 1), S134–S141, 1998.<br />

29. Gottlieb, R. A., Mitochondria: Ignition chamber for apoptosis, Mol. Genet. Metab., 68(2), 227–231,<br />

1999.<br />

30. Nieminen, A. L., Apoptosis and necrosis in health and disease: Role of mitochondria, Int. Rev.<br />

Cytol., 224, 29–55, 2003.<br />

31. Weber, T., Dalen, H., Andera, L., Negre-Salvayre, A., Auge, N., Sticha, M., Lloret, A. et al.,<br />

Mitochondria play a central role in apoptosis induced <strong>by</strong> alpha-tocopheryl succinate, an agent<br />

with antineoplastic activity: Comparison with receptor-mediated pro-apoptotic signaling, Biochemistry,<br />

42(14), 4277–4291, 2003.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 799<br />

32. Richter, C., Oxidative Stress, mitochondria, and apoptosis, Restor. Neurol. Neurosci., 12(2,3),<br />

59–62, 1998.<br />

33. Gulbins, E., Dreschers, S., and Bock, J., Role of mitochondria in apoptosis, Exp. Physiol., 88(Pt 1),<br />

85–90, 2003.<br />

34. Hockenbery, D. M., Giedt, C. D., O’Neill, J. W., Manion, M. K., and Banker, D. E., Mitochondria<br />

and apoptosis: New therapeutic targets, Adv. Cancer Res., 85, 203–242, 2002.<br />

35. Heiligtag, S. J., Bredehorst, R., and David, K. A., Key role of mitochondria in cerulenin-mediated<br />

apoptosis, Cell Death Differ., 9(9), 1017–1025, 2002.<br />

36. Boichuk, S. V., Minnebaev, M. M., and Mustafin, I. G., Key role of mitochondria in apoptosis of<br />

lymphocytes, Bull. Exp. Biol. Med., 132(6), 1166–1168, 2001.<br />

37. Birbes, H., Bawab, S. E., Obeid, L. M., and Hannun, Y. A., Mitochondria and ceramide: Intertwined<br />

roles in regulation of apoptosis, Adv. Enzyme Regul., 42, 113–129, 2002.<br />

38. Naoi, M., Maruyama, W., Akao, Y., and Yi, H., Mitochondria determine the survival and death in<br />

apoptosis <strong>by</strong> an endogenous neurotoxin, N-methyl(R)salsolinol, and neuroprotection <strong>by</strong> propargylamines,<br />

J. Neural Transm., 109(5,6), 607–621, 2002.<br />

39. Roucou, X., Antonsson, B., and Martinou, J. C., Involvement of mitochondria in apoptosis, Cardiol.<br />

Clin., 19(1), 45–55, 2001.<br />

40. Wang, X., The expanding role of mitochondria in apoptosis, Genes Dev., 15(22), 2922–2933,<br />

2001.<br />

41. Raha, S. and Robinson, B. H., Mitochondria, oxygen free radicals, and apoptosis, Am. J. Med.<br />

Genet., 106(1), 62–70, 2001.<br />

42. Waterhouse, N. J., Goldstein, J. C., Kluck, R. M., Newmeyer, D. D., and Green, D. R., The (Holey)<br />

study of mitochondria in apoptosis, Methods Cell Biol., 66, 365–391, 2001.<br />

43. Sordet, O., Rebe, C., Leroy, I., Bruey, J. M., Garrido, C., Miguet, C., Lizard, G., Plenchette, S.,<br />

Corcos, L., and Solary, E., Mitochondria-targeting drugs arsenic trioxide and lonidamine <strong>by</strong>pass the<br />

resistance of TPA-differentiated leukemic cells to apoptosis, Blood, 97(12), 3931–3940, 2001.<br />

44. Gottlieb, R. A., Mitochondria and apoptosis, Biol. Signals Recept., 10(3,4), 147–161, 2001.<br />

45. Shi, Y., A structural view of mitochondria-mediated apoptosis, Nat. Struct. Biol., 8(5), 394–401,<br />

2001.<br />

46. Gottlieb, R. A., Role of mitochondria in apoptosis, Crit. Rev. Eukaryot. Gene. Expr., 10(3,4),<br />

231–239, 2000.<br />

47. Jiang, Z. F., Zhao, Y., Hong, X., and Zhai, Z. H., Nuclear apoptosis induced <strong>by</strong> isolated mitochondria,<br />

Cell Res., 10(3), 221–232, 2000.<br />

48. Brenner, C. and Kroemer, G., Apoptosis. Mitochondria—the death signal integrators, Science,<br />

289(5482), 1150–1151, 2000.<br />

49. Desagher, S. and Martinou, J. C., Mitochondria as the central control point of apoptosis, Trends Cell<br />

Biol., 10(9), 369–377, 2000.<br />

50. Kroemer, G., Mitochondrial control of apoptosis: An overview, In Mitochondria and Cell Death,<br />

Brown, G. C., Nicholls, D. G., and Cooper, C. E., Eds., Princton University Press, Princeton, NJ,<br />

pp. 1–15, 1999.<br />

51. Fulda, S., Susin, S. A., Kroemer, G., and Debatin, K. M., Molecular ordering of apoptosis induced <strong>by</strong><br />

anticancer drugs in neuroblastoma cells, Cancer Res., 58(19), 4453–4460, 1998.<br />

52. Ferlini, C., Raspaglio, G., Mozzetti, S., Distefano, M., Filippetti, F., Martinelli, E., Ferrandina, G.,<br />

Gallo, D., Ranelletti, F. O., and Scambia, G., Bcl-2 down-regulation is a novel mechanism of<br />

paclitaxel resistance, Mol. Pharmacol., 64(1), 51–58, 2003.<br />

53. von Haefen, C., Wieder, T., Essmann, F., Schulze-Osthoff, K., Dorken, B., and Daniel, P. T.,<br />

Paclitaxel-induced apoptosis in BJAB cells proceeds via a death receptor-independent, caspases-<br />

3/-8-driven mitochondrial amplification loop, Oncogene, 22(15), 2236–2247, 2003.<br />

54. Andre, N., Carre, M., Brasseur, G., Pourroy, B., Kovacic, H., Briand, C., and Braguer, D., Paclitaxel<br />

targets mitochondria upstream of caspase activation in intact human neuroblastoma cells, FEBS<br />

Lett., 532(1–2), 256–260, 2002.<br />

55. Kidd, J. F., Pilkington, M. F., Schell, M. J., Fogarty, K. E., Skepper, J. N., <strong>Taylor</strong>, C. W., and Thorn, P.,<br />

Paclitaxel affects cytosolic calcium signals <strong>by</strong> opening the mitochondrial permeability transition pore,<br />

J. Biol. Chem., 277(8), 6504–6510, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


800<br />

Nanotechnology for Cancer Therapy<br />

56. Makin, G. W., Corfe, B. M., Griffiths, G. J., Thistlethwaite, A., Hickman, J. A., and Dive, C.,<br />

Damage-induced Bax N-terminal change, translocation to mitochondria and formation of Bax<br />

dimers/complexes occur regardless of cell fate, Embo J., 20(22), 6306–6315, 2001.<br />

57. Varbiro, G., Veres, B., Gallyas, F., and Sumegi, B., Direct effect of Taxol on free radical formation<br />

and mitochondrial permeability transition, Free Radic. Biol. Med., 31(4), 548–558, 2001.<br />

58. Andre, N., Braguer, D., Brasseur, G., Goncalves, A., Lemesle-Meunier, D., Guise, S., Jordan, M. A.,<br />

and Briand, C., Paclitaxel induces release of cytochrome c from mitochondria isolated from human<br />

neuroblastoma cells’, Cancer Res., 60(19), 5349–5353, 2000.<br />

59. Carre, M., Carles, G., Andre, N., Douillard, S., Ciccolini, J., Briand, C., and Braguer, D., Involvement<br />

of microtubules and mitochondria in the antagonism of arsenic trioxide on paclitaxel-induced<br />

apoptosis, Biochem. Pharmacol., 63(10), 1831–1842, 2002.<br />

60. Perkins, C. L., Fang, G., Kim, C. N., and Bhalla, K. N., The role of Apaf-1, caspase-9, and bid<br />

proteins in etoposide- or paclitaxel-induced mitochondrial events during apoptosis, Cancer Res.,<br />

60(6), 1645–1653, 2000.<br />

61. Itoh, M., Noutomi, T., Toyota, H., and Mizuguchi, J., Etoposide-mediated sensitization of squamous<br />

cell carcinoma cells to tumor necrosis factor-related apoptosis-inducing ligand (TRAIL)-induced<br />

loss in mitochondrial membrane potential, Oral Oncol., 39(3), 269–276, 2003.<br />

62. Custodio, J. B., Cardoso, C. M., Madeira, V. M., and Almeida, L. M., Mitochondrial<br />

permeability transition induced <strong>by</strong> the anticancer drug etoposide, Toxicol. In Vitro, 15(4,5),<br />

265–270, 2001.<br />

63. Fujino, M., Li, X. K., Kitazawa, Y., Guo, L., Kawasaki, M., Funeshima, N., Amano, T., and Suzuki, S.,<br />

Distinct pathways of apoptosis triggered <strong>by</strong> FTY720, etoposide, and anti-Fas antibody in human<br />

T-lymphoma cell line (Jurkat cells), J. Pharmacol. Exp. Ther., 300(3), 939–945, 2002.<br />

64. Robertson, J. D., Gogvadze, V., Zhivotovsky, B., and Orrenius, S., Distinct pathways for stimulation<br />

of cytochrome c release <strong>by</strong> etoposide, J. Biol. Chem., 275(42), 32438–32443, 2000.<br />

65. Marchetti, P., Zamzami, N., Joseph, B., Schraen-Maschke, S., Mereau-Richard, C., Costantini, P.,<br />

Metivier, D., Susin, S. A., Kroemer, G., and Formstecher, P., The novel retinoid 6-[3-(1-adamantyl)-<br />

4-hydroxyphenyl]-2-naphtalene carboxylic acid can trigger apoptosis through a mitochondrial<br />

pathway independent of the nucleus, Cancer Res., 59(24), 6257–6266, 1999.<br />

66. Murphy, M. P., Slip and leak in mitochondrial oxidative phosphorylation, Biochim. Biophys. Acta,<br />

977, 123–141, 1989.<br />

67. Murphy, M. P. and Smith, R. A., Drug delivery to mitochondria: The key to mitochondrial medicine,<br />

Adv. Drug Deliv. Rev., 41(2), 235–250, 2000.<br />

68. Modica-Napolitano, J. S. and Aprille, J. R., Delocalized lipophilic cations selectively target the<br />

mitochondria of carcinoma cells, Adv. Drug Deliv. Rev., 49(1,2), 63–70, 2001.<br />

69. Modica-Napolitano, J. S. and Aprille, J. R., Basis for the selective cytotoxicity of rhodamine 123,<br />

Cancer Res., 47(16), 4361–4365, 1987.<br />

70. Modica-Napolitano, J. S. and Singh, K. K., Mitochondria as targets for detection and treatment of<br />

cancer,Expert Rev. Mol. Med. 2002 April 11; 2002:1–19.<br />

71. Modica-Napolitano, J. S., Koya, K., Weisberg, E., Brunelli, B. T., Li, Y., and Chen, L. B., Selective<br />

damage to carcinoma mitochondria <strong>by</strong> the rhodacyanine MKT-077, Cancer Res., 56(3), 544–550,<br />

1996.<br />

72. Manetta, A., Emma, D., Gamboa, G., Liao, S., Berman, M., and DiSaia, P., Failure to enhance the<br />

in vivo killing of human ovarian carcinoma <strong>by</strong> sequential treatment with dequalinium chloride and<br />

tumor necrosis factor, Gynecol. Oncol., 50(1), 38–44, 1993.<br />

73. Weiss, M. J., Wong, J. R., Ha, C. S., Bleday, R., Salem, R. R., Steele, G. D., and Chen, L. B.,<br />

Dequalinium, a topical antimicrobial agent, displays anticarcinoma activity based on selective<br />

mitochondrial accumulation, Proc. Natl. Acad. Sci. USA, 84(15), 5444–5448, 1987.<br />

74. Christman, J. E., Miller, D. S., Coward, P., Smith, L. H., and Teng, N. N., Study of the selective<br />

cytotoxic properties of cationic, lipophilic mitochondrial-specific compounds in gynecologic malignancies,<br />

Gynecol. Oncol., 39(1), 72–79, 1990.<br />

75. Davis, S., Weiss, M. J., Wong, J. R., Lampidis, T. J., and Chen, L. B., Mitochondrial and plasma<br />

membrane potentials cause unusual accumulation and retention of rhodamine 123 <strong>by</strong> human breast<br />

adenocarcinoma-derived MCF-7 cells, J. Biol. Chem., 260(25), 13844–13850, 1985.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


DQAsomes as Mitochondria-Targeted Nanocarriers for Anti-Cancer Drugs 801<br />

76. Koya, K., Li, Y., Wang, H., Ukai, T., Tatsuta, N., Kawakami, M., Shishido, T., and Chen, L. B.,<br />

MKT-077, a novel rhodacyanine dye in clinical trials, exhibits anticarcinoma activity in preclinical<br />

studies based on selective mitochondrial accumulation, Cancer Res., 56(3), 538–543, 1996.<br />

77. Weisberg, E. L., Koya, K., Modica-Napolitano, J., Li, Y., and Chen, L. B., In vivo administration of<br />

MKT-077 causes partial yet reversible impairment of mitochondrial function, Cancer Res., 56(3),<br />

551–555, 1996.<br />

78. Chiba, Y., Kubota, T., Watanabe, M., Matsuzaki, S. W., Otani, Y., Teramoto, T., Matsumoto, Y.,<br />

Koya, K., and Kitajima, M., MKT-077, localized lipophilic cation: Antitumor activity against human<br />

tumor xenografts serially transplanted into nude mice, Anticancer Res., 18(2A), 1047–1052, 1998.<br />

79. Chiba, Y., Kubota, T., Watanabe, M., Otani, Y., Teramoto, T., Matsumoto, Y., Koya, K., and<br />

Kitajima, M., Selective antitumor activity of MKT-077, a delocalized lipophilic cation, on normal<br />

cells and cancer cells in vitro, J. Surg. Oncol., 69(2), 105–110, 1998.<br />

80. Petit, T., Izbicka, E., Lawrence, R. A., Nalin, C., Weitman, S. D., and Von Hoff, D. D., Activity of<br />

MKT 077, a rhodacyanine dye, against human tumor colony-forming units, Anticancer Drugs, 10(3),<br />

309–315, 1999.<br />

81. Propper, D. J., Braybrooke, J. P., <strong>Taylor</strong>, D. J., Lodi, R., Styles, P., Cramer, J. A., Collins, W. C.<br />

et al., Phase I trial of the selective mitochondrial toxin MKT077 in chemo-resistant solid tumours,<br />

Ann. Oncol., 10(8), 923–927, 1999.<br />

82. Adams, D. M., Ji, W., Barth, R. F., and Tjarks, W., Comparative in vitro evaluation of dequalinium<br />

B, a new boron carrier for neutron capture therapy (NCT), Anticancer Res., 20(5B), 3395–3402,<br />

2000.<br />

83. Weissig, V., Mogel, H. J., Wahab, M., and Lasch, J., Computer simulations of DQAsomes, Proceed.<br />

Intl. Symp. Control. Rel. Bioact. Mater., 25, 312, 1998.<br />

84. Weissig, V., Lasch, L., Erdos, G., Meyer, H. W., Rowe, T. C., and Hughes, J., DQAsomes: A novel<br />

potential drug and gene delivery system made from Dequalinium, Pharm. Res., 15(2), 334–337,<br />

1998.<br />

85. De Rosa, M., Gambacorta, A., and Gliozi, A., Structure, biosynthesis, and physicochemical properties<br />

of archaebacterial lipds, Microbiol. Rev., 50, 70–80, 1986.<br />

86. Gambacorta, A., Gliozi, A., and De Rosa, M., Archaeal lipids and their biotechnological applications,<br />

World J. Microbiol. Biotechnol., 11, 115–131, 1995.<br />

87. Weissig, V. and Torchilin, V. P., Mitochondriotropic cationic vesicles: A strategy towards mitochondrial<br />

gene therapy, Curr. Pharm. Biotechnol., 1(4), 325–346, 2000.<br />

88. Weissig, V. and Torchilin, V. P., Cationic single-chain bolaamphiphiles as new materials for application<br />

in medicine and biotechnology. In Materials research society meeting, Boston, MA., EE.5.4,<br />

p. 173, 1999.<br />

89. Weissig, V., Lizano, C., Ganellin, C. R., and Torchilin, V. P., DNA binding cationic bolasomes with<br />

delocalized charge center: A structure-activity relationship study, STP Pharma Sci., 11, 91–96, 2001.<br />

90. Lasch, J., Meye, A., Taubert, H., Koelsch, R., Mansa-ard, J., and Weissig, V., Dequalinium vesicles<br />

form stable complexes with plasmid DNA which are protected from DNase attack, Biol. Chem.,<br />

380(6), 647–652, 1999.<br />

91. Weissig, V., Lizano, C., and Torchilin, V. P., Selective DNA release from DQAsome/DNA<br />

complexes at mitochondria-like membranes, Drug Deliv., 7(1), 1–5, 2000.<br />

92. Weissig, V., D’Souza, G. G., and Torchilin, V. P., DQAsome/DNA complexes release DNA upon<br />

contact with isolated mouse liver mitochondria, J. Control Release, 75(3), 401–408, 2001.<br />

93. Ravagnan, L., Marzo, I., Costantini, P., Susin, S. A., Zamzami, N., Petit, P. X., and Hirsch, F.,<br />

Lonidamine triggers apoptosis via a direct, Bcl-2-inhibited effect on the mitochondrial permeability<br />

transition pore., Oncogene, 18(16), 2537–2546, 1999.<br />

94. Jacotot, E., Costantini, P., Laboureau, E., Zamzami, N., Susin, S. A., and Kroemer, G., Mitochondrial<br />

membrane permeabilization during the apoptotic process, Ann. N. Y. Acad. Sci., 887,<br />

18–30, 1999.<br />

95. van der Woude, I., Wagenaar, A., Meekel, A. A., ter Beest, M. B., Ruiters, M. H., Engberts, J. B., and<br />

Hoekstra, D., Novel pyridinium surfactants for efficient, nontoxic in vitro gene delivery, Proc. Natl.<br />

Acad. Sci. USA, 94(4), 1160–1165, 1997.<br />

96. Zuhorn, I. S. and Hoekstra, D., On the mechanism of cationic amphiphile-mediated transfection.<br />

To fuse or not to fuse: Is that the question? J. Membr. Biol., 189(3), 167–179, 2002.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


802<br />

Nanotechnology for Cancer Therapy<br />

97. Wattiaux, R., Jadot, M., Warnier-Pirotte, M. T., and Wattiaux-De Coninck, S., Cationic lipids<br />

destabilize lysosomal membrane in vitro, FEBS Lett., 417(2), 199–202, 1997.<br />

98. El Ouahabi, A., Thiry, M., Pector, V., Fuks, R., Ruysschaert, J. M., and Vandenbranden, M., The role<br />

of endosome destabilizing activity in the gene transfer process mediated <strong>by</strong> cationic lipids, FEBS<br />

Lett., 414(2), 187–192, 1997.<br />

99. D’Souza, G. G., Rammohan, R., Cheng, S. M., Torchilin, V. P., and Weissig, V., DQAsomemediated<br />

delivery of plasmid DNA toward mitochondria in living cells, J. Control Release, 92,<br />

189–197, 2003.<br />

100. D’Souza, G. G., Boddapati, S. V., and Weissig, V., Mitochondrial leader sequence-plasmid DNA<br />

conjugates delivered into mammalian cells <strong>by</strong> DQAsomes co-localize with mitochondria, Mitochondrion,<br />

5(5), 352–358, 2005.<br />

101. Braguer, D., Andre, N., Carre, M., Carles, G., Gonzalves, A., and Briand, C., 92nd Annual Meeting of<br />

the American Association for Cancer Research, American Association for Cancer Research, New<br />

Orleans, LA, p. 369, 2001.<br />

102. Carre, M., Andre, N., Carles, G., Borghi, H., Brichese, L., Briand, C., and Braguer, D., Tubulin is an<br />

inherent component of mitochondrial membranes that interacts with the voltage-dependent anion<br />

channel, J. Biol. Chem., 277, 33664–33669, 2002.<br />

103. Cheng, S. M., Pabba, S., Torchilin, V. P., Fowle, W., Kimpfler, A., Schubert, R., and Weissig, V.,<br />

Towards mitochondria-specific delivery of apoptosis-inducing agents: DQAsomal incorporated<br />

paclitaxel, J. Drug Deliv. Sci. Techno., 15(1), 81–86, 2005.<br />

104. Discher, E. D. and Eisenberg, A., Polymer vesicles, Science, 297, 967–973, 2002.<br />

105. Schuppenhauer, M., Munich Biotech: Catioms for cancer. BioCentury, The Bernstein Report on<br />

BioBusiness, May 19, A10, 2003.<br />

106. Schmitt-Sody, M., Strieth, S., Krasnici, S., Sauer, B., Schulze, B., Teifel, M., Michaelis, U., Naujoks, K.,<br />

and Dellian, M., Neovascular targeting therapy: Paclitaxel encapsulated in cationic liposomes improves<br />

antitumoral efficacy, Clinical Cancer Research, 9(6), 2335–2341, 2003.<br />

107. Krasnici, S., Werner, A., Eichhorn, M. E., Schmitt-Sody, M., Pahernik, S. A., Sauer, B., and Schulze, B.,<br />

Effect of the surface charge of liposomes on their uptake <strong>by</strong> angiogenic tumor vessels, Int. J. Cancer,<br />

105(4), 561–567, 2003.<br />

108. Thurston, G., McLean, J. W., Rizen, M., Baluk, P., Haskell, A., Murphy, T. J., Hanahan, D., and<br />

McDonald, D. M., Cationic liposomes target angiogenic endothelial cells in tumors and chronic<br />

inflammation in mice, J. Clin. Invest., 101(7), 1401–1413, 1998.<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong>


Section 1<br />

Nanotechnology and Cancer<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

1


Section 2<br />

Polymer Conjugates<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

157


Section 3<br />

Polymeric Nanoparticles<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

213


Section 4<br />

Polymeric Micelles<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

315


Section 5<br />

Dendritic Nanocarriers<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

487


Section 6<br />

Liposomes<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

593


Section 7<br />

Other Lipid Nanostructures<br />

q <strong>2006</strong> <strong>by</strong> <strong>Taylor</strong> & <strong>Francis</strong> <strong>Group</strong>, <strong>LLC</strong><br />

721

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!